lagrangian and dirac constraints for the ideal ...morrison/20_mapiic_arx.pdfsitael s.p.a., pisa,...

54
Lagrangian and Dirac constraints for the ideal incompressible fluid and magnetohydrodynamics P. J. Morrison * Department of Physics and Institute for Fusion Studies, The University of Texas at Austin, Austin, TX 78712-1060, USA T. Andreussi SITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract The incompressibility constraint for fluid flow was imposed by Lagrange in the so-called Lagrangian vari- able description using his method of multipliers in the Lagrangian (variational) formulation. An alternative is the imposition of incompressibility in the Eulerian variable description by a generalization of Dirac’s con- straint method using noncanonical Poisson brackets. Here it is shown how to impose the incompressibility constraint using Dirac’s method in terms of both the canonical Poisson brackets in the Lagrangian variable description and the noncanonical Poisson brackets in the Eulerian description, allowing for the advection of density. Both cases give dynamics of infinite-dimensional geodesic flow on the group of volume preserving dieomorphisms and explicit expressions for this dynamics in terms of the constraints and original variables is given. Because Lagrangian and Eulerian conservation laws are not identical, comparison of the various methods is made. * [email protected] 1

Upload: others

Post on 21-Sep-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Lagrangian and Dirac constraints for the ideal incompressible fluid and

magnetohydrodynamics

P. J. Morrison∗

Department of Physics and Institute for Fusion Studies,

The University of Texas at Austin, Austin, TX 78712-1060, USA

T. Andreussi

SITAEL S.p.A., Pisa, 56121, Italy

F. Pegoraro

Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy

Abstract

The incompressibility constraint for fluid flow was imposed by Lagrange in the so-called Lagrangian vari-

able description using his method of multipliers in the Lagrangian (variational) formulation. An alternative

is the imposition of incompressibility in the Eulerian variable description by a generalization of Dirac’s con-

straint method using noncanonical Poisson brackets. Here it is shown how to impose the incompressibility

constraint using Dirac’s method in terms of both the canonical Poisson brackets in the Lagrangian variable

description and the noncanonical Poisson brackets in the Eulerian description, allowing for the advection of

density. Both cases give dynamics of infinite-dimensional geodesic flow on the group of volume preserving

diffeomorphisms and explicit expressions for this dynamics in terms of the constraints and original variables

is given. Because Lagrangian and Eulerian conservation laws are not identical, comparison of the various

methods is made.

[email protected]

1

Page 2: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

I. INTRODUCTION

A. Background

Sometimes constraints are maintained effortlessly, an example being ∇ · B = 0 in electrody-

namics which if initially true remains true, while alternatively in most cases dynamical equations

must be modified to maintain constraints, an example being ∇·v = 0 in fluid mechanics. The need

to apply constraints arises in a variety of contexts, ranging from gauge field theories [e.g. 83] to

optimization and control [e.g. 54]. A very common approach is to use the method of Lagrange

multipliers, which is taught in standard physics curricula for imposing holonomic constraints in

mechanical systems. Alternatively, Dirac [62], in pursuit of his goal of quantizing gauge field

theories, introduced a method that uses the Poisson bracket.

The purpose of the present article is to explore different methods for imposing the compress-

ibility constraint in ideal (dissipation-free) fluid mechanics and its extension to magnetohydrody-

namics (MHD), classical field theories intended for classical purposes. This endeavor is richer

than might be expected because the different methods of constraint can be applied to the fluid in

either the Lagrangian (material) description or the Eulerian (spatial) description, and the constraint

methods have different manifestations in the Lagrangian (action principle) and Hamiltonian for-

mulations. Although Lagrange’s multiplier is widely appreciated, it is not so well known that he

used it long ago for imposing the incompressibility constraint for a fluid in the Lagrangian vari-

able description [69]. More recently, Dirac’s method was applied for imposing incompressibility

within the Eulerian variable description, first in Nguyen & Turski [77, 78] and followed up in

several works [55, 56, 57, 75, 84]. Given that a Lagrangian conservation law is not equivalent to

an Eulerian conservation law, it remains to elucidate the interplay between the methods of con-

straint and the variables used for the description of the fluid. Thus we have three dichotomies: the

Lagrangian vs. Eulerian fluid descriptions, Lagrange multiplier vs. Dirac constraint methods, and

Lagrangian vs. Hamiltonian formalisms. It is the elucidation of the interplay between these, along

with generalizing previous results, that is the present goal.

It is well known that a free particle with holonomic constraints, imposed by the method of

Lagrange multipliers, is a geodesic flow. Indeed, Lagrange essentially observed this in Lagrange

[69] for the ideal fluid when he imposed the incompressibility constraint by his method. Lagrange

did this in the Lagrangian description by imposing the constraint that the map from initial positions

2

Page 3: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

of fluid elements to their positions at time t preserves volume, and he did this by the method of

Lagrange multipliers. It is worth noting that Lagrange knew the Lagrange multiplier turns out to be

the pressure, but he had trouble solving for it. Lagrange’s calculation was placed in a geometrical

setting by Arnold [50] (see also Appendix 2 of [51]), where the constrained maps from the initial

conditions were first referred to as volume preserving diffeomorphisms in this context. Given this

background, in our investigation of the three dichotomies described above we emphasize geodesic

flow.

For later use we record here the incompressible Euler equations for the case with constant

density and the case where density is advected. The equations of motion, allowing for density

advection, are given by

∂v∂t

= −v · ∇v −1ρ∇p , (1)

∇ · v = 0 , (2)

∂ρ

∂t= −v · ∇ρ , (3)

where v(x, t) is the velocity field, ρ(x, t) is the mass density, p(x, t) is the pressure, and x ∈ D,

the region occupied by the fluid. These equations are generally subject to the free-slip boundary

condition n · v|∂D = 0, where n is normal to the boundary of D. The pressure field that enforces

the constraint (2) is obtained by setting ∂(∇ · v)∂t = 0, which implies

∆ρp := ∇ ·(1ρ∇p

)= −∇ · (v · ∇v) . (4)

For reasonable assumptions on ρ and boundary conditions, (4) is a well-posed elliptic equation

[see e.g. 63], so we can write

p = −∆−1ρ ∇ · (v · ∇v) . (5)

For the case where ρ is constant we have the usual Green’s function expression

p(x, t) = − ρ

∫d3x′G(x, x′)∇′ · (v′ · ∇′v′) , (6)

where G is consistent with Neumann boundary conditions [79] and v′ = v(x′, t). Insertion of (6)

into (1) gives∂v∂t

= −v · ∇v + ∇

∫d3x′G(x, x′)∇′ · (v′ · ∇′v′) , (7)

which is a closed system for v(x, t).

For MHD, equation (1) has the additional term (∇ × B) × B/ρ added to the righthand side.

Consequently for this model, the source of (5) is modified by the addition of this term to −v · ∇v.

3

Page 4: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

B. Overview

Section II contains material that serves as a guide for navigating the more complicated calcu-

lations to follow. We first consider the various approaches to constraints in the finite-dimensional

context in Sections II A–II C. Section II A briefly covers conventional material about holonomic

constraints by Lagrange multipliers – here the reader is reminded how the free particle with holo-

nomic constraints amounts to geodesic flow. Section II B begins with the phase space action

principle, whence the Dirac bracket for constraints is obtained by Lagrange’s multiplier method,

but with phase space constraints as opposed to the usual holonomic configuration space constraints

used in conjunction with Hamilton’s principle of mechanics, as described in Section II A. Next, in

Section II C, we compare the results of Sections II A and II C and show how conventional holo-

nomic constraints can be enforced by Dirac’s method. Contrary to Lagrange’s method, here we

obtain explicit expressions, ones that do not appear in conventional treatments, for the Christoffel

symbol and the normal force entirely in terms of the original Euclidean coordinates and con-

straints. Section II is completed with Section II D, where the previous ideas are revisited in the

d+1 field theory context in prepartion for the fluid and MHD calculations. Holonomic constraints,

Dirac brackets, with local or nonlocal constraints, and geodesic flow are treated.

In Section III we first consider the compressible (unconstrained) fluid and MHD versions of

Hamilton’s variational principle, the principle of least action, with Lagrange’s Lagrangian in the

Lagrangian description. From this we obtain in Section III B the canonical Hamiltonian field

theoretic form in the Lagrangian variable description, which is transformed in Section III C, via

the mapping from Lagrangian to Eulerian variables, to the noncanonical Eulerian form. Section

III is completed by an in depth comparison of constants of motion in the Eulerian and Lagrangian

descriptions, which surprisingly does not seem to appear in fluid mechanics or plasma physics

textbooks. The material of this section is necessary for understanding the different manifestations

of constraints in our dichotomies.

Section IV begins with Section IV A that reviews Lagrange’s original calculations. Because

the incompressibility constraint he imposes is holonomic and there are no additional forces, his

equations describe infinite-dimensional geodesics flow on volume preserving maps. The remain-

ing portion of this section contains the most substantial calculations of the paper. In Section IV B

for the first time Dirac’s theory is applied to enforce incompressibility in the Lagrangian variable

description. This results in a new Dirac bracket that generates volume preserving flows. As in Sec-

4

Page 5: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

tion II C 1, which serves as a guide, the equations of motion generated by the bracket are explicit

and contain only the constraints and original variables. Next, in Section IV C, a reduction from

Lagrangian to Eulerian variables is made, resulting in a new Eulerian variable Poisson bracket that

allows for density advection while preserving incompressibility. This was an heretofore unsolved

problem. Section IV D ties together the results of Sections IV B, IV C, and III D. Here both the

Eulerian and Lagrangian Dirac constraint theories are compared after they are evaluated on their

respective constraints, simplifying their equations of motion. Because Lagrangian and Eulerian

conservation laws are not identical, we see that there are differences. Section IV concludes in

Section IV E with a discussion of the full algebra of invariants, that of the ten parameter Galilean

group, for both the Lagrangian and Eulerian descriptions. In addition the Casimir invariants of the

theories are discussed.

The paper concludes with Section V, where we briefly summarize our results and speculate

about future possibilities.

II. CONSTRAINT METHODS

A. Holonomic constraints by Lagrange’s multiplier method

Of interest are systems with Lagrangians of the form L(q, q) where the overdot denotes time

differentiation and q = (q1, q2, . . . , qN). Because nonautonomous systems could be included by

appending an additional degree of freedom, explicit time dependence is not included in L.

Given the Lagrangian, the equations of motion are obtained according to Hamilton’s principle

by variation of the action

S [q] =

∫ t1

t0dt L(q, q) ; (8)

i.e.

δS [q; η] :=ddε

S [q + εδq]∣∣∣∣ε=0

=

∫ t1

t0dt

(ddt∂L∂qi −

∂L∂qi

)δqi =

∫ t1

t0dtδS [q]δqi(t)

δqi = 0 , (9)

for all variations δq(t) satisfying δq(t0) = δq(t1) = 0, implies Lagrange’s equations of motion, i.e.

δS [q]δqi(t)

= 0 ⇒ddt∂L∂qi −

∂L∂qi = 0 , i = 1, 2, . . . ,N . (10)

Holonomic constraints are real-valued functions of the form CA(q), A = 1, 2, . . . ,M, which are

desired to be constant on trajectories. Lagrange’s method for implementing such constraints is to

5

Page 6: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

add them to the action and vary as follows:

δS λ := δ(S + λACA) = 0 , (11)

yielding the equations of motion

δS λ[q]δqi(t)

= 0 ⇒ddt∂L∂qi −

∂L∂qi = λA

∂CA

∂qi , i = 1, 2, . . . ,N , (12)

with the forces of constraint residing on the right-hand side of (12). Observe in (11) and (12)

repeated sum notation is implied for the index A. The N equations of (12) with the M numerical

values of the constraints CA(q) = CA0 , determine the N + M unknowns {qi} and {λA}. In practice,

because solving for the Lagrange multipliers can be difficult an alternative procedure, a example

of which we describe in Sec. II A 1, is used.

We will see in Section IV A that the field theoretic version of this method is how Lagrange

implemented the incompressiblity constraint for fluid flow. For the purpose of illustration and in

preparation for later development, we consider a finite-dimensional analog of Lagrange’s treat-

ment.

1. Holonomic constraints and geodesic flow via Lagrange

Consider N noninteracting bodies each of mass mi in the Eucledian configuration space E3N

with cartesian coordinates qi = (qxi, qyi, qzi), where i = 1, 2, . . . ,N. The Lagrangian for this system

is given by the usual kinetic energy,

L =

N∑i=1

mi

2qi · qi , (13)

with the usual “dot” product. The Euler-Lagrange equations for this system, equations (10), are

the uninteresting system of N free particles. As in Sec. II A we constrain this system by adding

constraints CA(q1,q2, . . . ,qN), where again A = 1, 2, . . . ,M, leading to the equations

miqi = λA∂CA

∂qi. (14)

Instead of solving the 3N equations of (14) together with the M numerical values of the con-

straints, in order to determine the unknowns qi and λA, we recall the alternative procedure, which

dates back to Lagrange (see e.g. Sec. IV of Lagrange [69]) and has been taught to physics students

6

Page 7: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

for generations [see e.g. 59, 86]. With the alternative procedure one introduces generalized coor-

dinates that account for the constraints, yielding a smaller system on the constraint manifold, one

with the Lagrangian

L =12

gµν(q) qµqν , µ, ν = 1, 2 . . . , 3N − M , (15)

where

gµν =

N∑i=1

mi∂qi

∂qµ·∂qi

∂qν. (16)

Then Lagrange’s equations (10) for the Lagrangian (15) are the usual equations for geodesic flow

qµ + Γµαβ qαqβ = 0 , (17)

where the Christoffel symbol is as usual

Γµαβ =

12

gµν(

gνα∂qβ

+gνβ∂qα−

gαβ∂qν

). (18)

If the constraints had time dependence, then the procedure would have produced the Coriolis and

centripetal forces, as is usually done in textbooks.

Thus, we arrive at the conclusion that free particle dynamics with time-independent holonomic

constraints is geodesic flow.

B. Dirac’s bracket method

So, a natural question to ask is “How does one implement constraints in the Hamiltonian set-

ting, where phase space constraints depend on both the configuration space coordinate q and the

canonical momentum p”? [See e.g. 53, 83]. To this end we begin with the phase space action

principle

S [q, p] =

∫ t1

t0dt

[piqi − H(q, p)

], (19)

where again repeated sum notation is used for i = 1, 2, . . . ,N. Independent variation of S [q, p]

with respect to q and p, with δq(t0) = δq(t1) = 0 and no conditions on δp, yields Hamilton’s

equations,

pi = −∂H∂qi and qi =

∂H∂pi

, (20)

or equivalently

zα = {zα,H} , (21)

7

Page 8: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

which is a rewrite of (20) in terms of the Poisson bracket on phase space functions f and g,

{ f , g} =∂ f∂qi

∂g∂pi−∂g∂qi

∂ f∂pi

=∂ f∂zαJαβc

∂g∂zβ

, (22)

where in the second equality we have used z = (q, p), so α, β = 1, 2, . . . , 2N and the cosymplectic

form (Poisson matrix) is

Jc =

ON IN

−IN ON

, (23)

with ON being an N × N block of zeros and IN being the N × N identity.

Proceeding as in Section II A, albeit with phase space constraints Da(q, p), a = 1, 2, . . . , 2M <

2N, we vary

S λ[q, p] =

∫ t1

t0dt

[piqi − H(q, p) + λaDa

], (24)

and obtain

pi = −∂H∂qi + λa

∂Da

∂qi and qi =∂H∂pi− λa

∂Da

∂pi. (25)

Next, enforcing Da = 0 for all a, will ensure that the constraints stay put. Whence, differentiating

the Da and using (25) yields

Da =∂Da

∂qi qi +∂Da

∂pipi

= {Da,H} − λb{Da,Db} ≡ 0 . (26)

We assume Dab := {Da,Db} has an inverse, D−1ab , which requires there be an even number of

constraints, a, b = 1, 2, . . . , 2M, because odd antisymmetric matrices have zero determinant. Then

upon solving (26) for λb and inserting the result into (25) gives

zα = {zα,H} − D−1ab {z

α,Da}{Db,H} . (27)

From (27), we obtain a generalization of the Poisson bracket, the Dirac bracket,

{ f , g}∗ = { f , g} − D−1ab { f ,D

a}{Db, g} . (28)

which has the degeneracy property

{ f ,Da}∗ ≡ 0 . (29)

for all functions f and indices a = 1, 2, . . . , 2M.

The generation of the equations of motion via a Dirac bracket, i.e.

zα = {zα,H}∗ , (30)

8

Page 9: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

which is equivalent to (27), has the advantage that the Lagrange multipliers λA have been elimi-

nated from the theory.

Note, although the above construction of the Dirac bracket started with the canonical bracket

of (III C), his construction results in a valid Poisson bracket if one starts from any valid Poisson

bracket (cf. (83) of Section II D and Section III C), which need not have a Poisson matrix of the

form of (23) [see e.g. 75]. We will use such a bracket in Section IV C when we apply constraints

by Dirac’s method in the Eulerian variable picture. Also note, for our purposes it is not necessary

to describe primary vs. secondary constraints (although we use the latter), and the notions of weak

vs. strong equality. We refer the reader to Dirac [62], Hanson et al. [66], Sudarshan & Makunda

[82], Sundermeyer [83] for treatment of these concepts.

C. Holonomic constraints by Dirac’s bracket method

A connection between the approaches of Lagrange and Dirac can be made. From a set of La-

grangian constraints CA(q), where A = 1, 2, . . . ,M, one can construct an additional M constraints

by differentiation,

CA =∂CA

∂qi qi =∂CA

∂qi

∂H∂pi

, (31)

where the second equality is possible if (10) possesses the Legendre transformation to the Hamil-

tonian form. In this way we obtain an even number of constraints

Da(q, p) =(CA(q) , CA′(q, p)

), (32)

where A = 1, 2, . . . ,M, A′ = M + 1,M + 2, . . . , 2M, and a = 1, 2, . . . , 2M.

With the constraints of (32) the bracket Dab = {Da,Db} needed to construct (28) is easily

obtained,

D =

OM {CA, CB′}

{CA′ ,CB} {CA′ , CB′}

=:

OM S

−S A

, (33)

where OM is an M × M block of zeros and S is the following M × M symmetric matrix with

elements

SAB := {CA, CB} =∂2H∂pi∂p j

∂CA

∂qi

∂CB

∂q j , (34)

9

Page 10: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

and A is the following M × M antisymmetric matrix with elements

AAB := {CA′ , CB′} (35)

=∂2H∂pi∂pk

[∂2H∂qi∂p j

(∂CA

∂q j

∂CB

∂qk −∂CB

∂q j

∂CA

∂qk

)+∂H∂p j

(∂2CA

∂qi∂q j

∂CB

∂qk −∂2CB

∂qi∂q j

∂CA

∂qk

)].

Assuming the existence of D−1, the 2M × 2M inverse of (33), the Dirac bracket of (28) can be

constructed. A necessary and sufficient condition for the existence of this inverse is that detS , 0,

and when this is the case the inverse is given by

D−1 =

S−1AS−1 −S−1

S−1 OM

. (36)

Because of the block diagonal structure of (36), the Dirac bracket (28) becomes

{ f , g}∗ = { f , g} + S−1AB

({ f ,CA}{CB, g} − {g,CA}{CB, f }

)+ S−1

AC ACD S−1

DB { f ,CA}{CB, g} , (37)

which has the form

{ f , g}∗ = { f , g} − (P⊥)ij

(∂ f∂qi

∂g∂p j−∂g∂qi

∂ f∂p j

)+ Qi j ∂ f

∂pi

∂g∂p j

=∂ f∂qi P

ij∂g∂p j−∂g∂qi P

ij∂ f∂p j

+ Qi j ∂ f∂pi

∂g∂p j

, (38)

where the matrices P = I − P⊥, with

(P⊥)ij = S−1

AB∂2H∂pi∂pk

∂CA

∂q j

∂CB

∂qk , (39)

and Q, a complicated expression that we will not record, are crafted using the constraints and

Hamiltonian so as to make { f , g}∗ preserve the constraints.

The equations of motion that follow from (37) are

q` = {q`,H}∗ =∂H∂p`

+ S−1AB

({q`,CA}{CB,H} − {H,CA}{CB, q`}

)+ S−1

AC ACD S−1

DB {q`,CA}{CB,H} , (40)

p` = {p`,H}∗ = −∂H∂q`

+ S−1AB

({p`,CA}{CB,H} − {H,CA}{CB, p`}

)+ S−1

AC ACD S−1

DB {p`,CA}{CB,H} . (41)

10

Page 11: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Given the Dirac bracket associated with the D of (34), dynamics that enforces the constraints

takes the form of (30). Any system generated by this bracket will enforce Lagrange’s holonomic

constraints; however, only initial conditions compatible with

Da ≡ 0 , ∀ a = M + 1,M + 2, . . . , 2M , (42)

or equivalently

CA =∂CA

∂qi

∂H∂pi

= {CA,H} ≡ 0 , ∀ A = 1, 2, . . . ,M , (43)

will correspond to the system with holonomic constraints. Using (43) and {q`,CA} ≡ 0, (40) and

(41) reduce to

q` = {q`,H}∗ =∂H∂p`

, (44)

p` = {p`,H}∗ = −∂H∂q`

+ S−1AB{p`,C

A}{CB,H} , (45)

where

{CB,H} =

(∂2H∂qi∂p j

∂CB

∂q j +∂H∂p j

∂2CB

∂qiq j

)∂H∂pi−∂CB

∂qi

∂2H∂pi∂p j

∂H∂q j . (46)

Thus the Dirac bracket approach gives a relatively simple system for enforcing holonomic con-

straints. It can be shown directly that if initially CA vanishes, then the system of (44) and (45) will

keep it so for all time.

1. Holonomic constraints and geodesic flow via Dirac

Let us now consider again the geodesic flow problem of Sec. II A 1: the N degree-of-freedom

free particle system with holonomic constraints, but this time within the framework of Dirac

bracket theory. For this problem the unconstrained configuration space is the Euclidean space

E3N and we will denote by Q the constraint manifold within E3N defined by the constancy of the

constraints CA.

The Lagrangian of (13) is easily Legendre transformed to the free particle Hamiltonian

H =

N∑i=1

12mi

pi · pi , (47)

where pi = miqi. For this example the constraints of (32) take the form

Da =

(CA(q1,q2, . . . ,qN),

∂CA′(q1,q2, . . . ,qN)∂qi

·pi

mi

), (48)

11

Page 12: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

the M × M matrix S has elements

SAB =

N∑i=1

1mi

∂CA

∂qi·∂CB

∂qi, (49)

and the M × M matrix A is

AAB =

N∑i, j=1

1mim j

pi ·

[∂2CA

∂qi∂q j·∂CB

∂q j−

∂2CB

∂qi∂q j·∂CA

∂q j

]. (50)

The Dirac bracket analogous to (38) is

{ f , g}∗ =

N∑i j=1

[∂ f∂qi·↔

Pi j ·∂g∂p j−∂g∂qi·↔

Pi j ·∂ f∂p j

+∂ f∂pi·↔

Qi j ·∂g∂p j

], (51)

where↔

Pi j =↔

I i j −↔

P⊥ i j with the tensors

P⊥ i j :=N∑

k=1

S−1AB

∂2H∂pi∂pk

·∂CB

∂qk

∂CA

∂q j= S−1

AB1mi

∂CB

∂qi

∂CA

∂q j, (52)

Q i j :=N∑

k=1

S−1AB

[∂CA

∂q j

pk

mk·∂2CB

∂qk∂qi−∂CA

∂qi

pk

mk·∂2CB

∂qk∂q j

]+ S−1

AC ACD S−1

DB∂CA

∂qi

∂CB

∂q j(53)

=:↔

Ti j −↔

T ji +↔

Ai j , (54)

where↔

Ai j is the term with S−1ACA

CDS−1DB. Observe

Ai j = −↔

A ji because ACD = −ADC and

N∑k=1

P⊥ ik ·↔

P⊥ k j =

N∑k=1

(S−1

AB1mi

∂CB

∂qi

∂CA

∂qk

(S−1

A′B′1

mk

∂CB′

∂qk

∂CA′

∂q j

)

=

(S−1

AB1mi

∂CB

∂qi

)S−1

A′B′

N∑k=1

∂CA

∂qk·

1mk

∂CB′

∂qk

∂CA′

∂q j

=

(S−1

AB1mi

∂CB

∂qi

)S−1

A′B′ SAB′ ∂CA′

∂q j=↔

P⊥ i j . (55)

Also observe for the Hamiltonian of (47)

N∑j=1

P⊥ i j ·∂H∂p j

=

N∑j=1

P⊥ i j ·p j

m j≡ 0 , (56)

N∑j=1

∂H∂p j·↔

Ti j =

N∑j=1

p j

m j·↔

Ti j ≡ 0 , (57)

N∑j=1

Ai j ·∂H∂p j

= −

N∑j=1

A ji ·∂H∂p j≡ 0 , (58)

12

Page 13: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

when evaluated on the constraint CA,B = 0, while

N∑j=1

∂H∂p j·↔

P⊥ ji , 0 andN∑

i=1

∂H∂pi·↔

Ti j , 0 , (59)

when evaluated on the constraint CA,B = 0. Thus, the bracket of (51) yields the equations of motion

qi = {qi,H}∗ =∂H∂pi

=pi

mi, (60)

pi = −∂CA

∂qiS−1

AB

N∑j,k=1

p j

m j·∂2CB

∂q jqk·

pk

mk, (61)

or

qi = −1mi

∂CA

∂qiS−1

AB

N∑j,k=1

q j ·∂2CB

∂q jqk· qk = −

N∑j,k=1

q j · Γi, jk · qk , (62)

where

Γi, jk :=1mi

∂CA

∂qi⊗ S−1

AB∂2CB

∂q jqk, (63)

is used to represent the normal force.

Observe, (62) has two essential features: as noted, its righthand side is a normal force that

projects to the constraint manifold defined by the constraints CA and within the constraint manifold

it describes a geodesic flow, all done in terms of the original Euclidean space coordinates where

the initial conditions place the flow on Q by setting the values CA for all A = 1, 2, . . . ,M. We will

show this explicitly.

First, because the components of vectors normal toQ are given by ∂CA/∂qi for A = 1, 2, . . . ,M,

this prefactor on the righthand side of (62) projects as expected. Upon comparing (62) with (14)

we conclude that the coefficient of this prefactor must be the Lagrange multipliers, i.e.,

λA = −S−1AB

N∑k, j=1

q j ·∂2CB

∂q jqk· qk . (64)

Thus, we see that Dirac’s procedure explicitly solves for the Lagrange multiplier.

Second, to uncover the geodesic flow we can proceed as usual by projecting explicitly onto

Q. To this end we consider the transformation between the Euclidean configuration space E3N

coordinates

(q1,q2, . . . ,qi, . . . ,qN) , where i = 1, 2, . . . ,N (65)

and another set of coordinates

(q1, q2, . . . , qa, . . . , q3N) , where a = 1, 2, . . . , 3N , (66)

13

Page 14: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

which we tailor as follows:

(q1, q2, . . . , qα . . . , qn,C1,C2, . . . ,CA, . . . ,CM) , (67)

where α = 1, 2, . . . , n, A = 1, 2, . . . ,M, and n = 3N − M. Here we have chosen qn+A = CA and n

is the actual number of degrees of freedom on the constraint surface Q. We can freely transform

back and forth between the two coordinates, i.e.

(q1,q2, . . . ,qi, . . . ,qN)←→ (q1, q2, . . . , qa, . . . , q3N) . (68)

Note, the choice qn+A = CA could be replaced by qn+A = f A(C1,C2, . . . ,CM) for arbitrary indepen-

dent f A, but we assume the original CA are optimal. Because qα are coordinates within Q, tangent

vectors to Q have the components ∂qi/∂qα, and there is one for each α = 1, 2, . . . , n. The pairing

of the normals with tangents is expressed by

N∑i=1

∂qi

∂qα·∂CA

∂qi= 0 , α = 1, 2, . . . , n; A = 1, 2, . . . ,M . (69)

Let us now consider an alternative procedure that the Dirac constraint method provides. Proceed-

ing directly we calculate

qa =

N∑i=1

∂qa

∂qi· qi . (70)

Observe that on E3N the matrix ∂qa/∂qi is invertible and the full metric tensor and its inverse in

the new coordinates are given as follows:

gab =

N∑i=1

1mi

∂qa

∂qi·∂qb

∂qiand gab =

N∑i=1

mi∂qi

∂qa ·∂qi

∂qb . (71)

The metric tensor on Q of (16) is obtained by restricting gab to a, b ≤ n and gαβ is obtained by

inverting gαβ and not by restricting gab. Proceeding by differentiating again we obtain

qa =

N∑i=1

∂qa

∂qi· qi +

N∑i, j=1

qi ·∂2qa

∂qi∂q j· q j , a = 1, 2, . . . , 3N. (72)

Now inserting (62) into (72) gives

qa = −

N∑i=1

1mi

∂qa

∂qi·∂CA

∂qigAB

N∑j,k=1

q j ·∂2CB

∂q jqk· qk +

N∑i, j=1

q j ·∂2qa

∂qi∂q j· qi , (73)

where we have recognized that

gAB = SAB =

N∑i=1

1mi

∂CA

∂qi·∂CB

∂qi(74)

14

Page 15: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

and, as was necessary for the workability of the Dirac bracket constraint theory, gAB = S−1AB must

exist. This quantity is obtained by inverting SAB and not by restricting gab.

Equation (73) is an expression for the full system on E3N . However, for a > n, we know

qa = CA = 0, so the two terms of (73) should cancel. To see this, in the first term of (73) we set

qa = CC and observe that this first term becomes

N∑i=1

1mi

∂CC

∂qi·∂CA

∂qigAB

N∑j,k=1

q j ·∂2CB

∂q jqk· qk = −gCA gAB

N∑j,k=1

q j ·∂2CB

∂q jqk· qk

= −

N∑j,k=1

q j ·∂2CC

∂q jqk· qk . (75)

Now, for a ≤ n, say α, the righthand side gives a Christoffel symbol expression for the geodesic

flow; viz.

qα = −

N∑i=1

1mi

∂qα

∂qi·∂CA

∂qigAB

N∑j,k=1

q j ·∂2CB

∂q jqk· qk +

N∑j,k=1

q j ·∂2qα

∂q j∂qk· qk

= −Γαµν qµqν , (76)

where

Γαµν =

N∑i=1

1mi

∂qα

∂qi·∂CA

∂qiS−1

AB

N∑j,k=1

∂q j

∂qµ·∂2CB

∂q jqk·∂qk

∂qν+

N∑j,k=1

∂q j

∂qµ·∂2qα

∂q j∂qk·∂qk

∂qν(77)

is an expression for the Christoffel symbol in terms of the original Euclidean coordinates, the

constraints, and the choice of coordinates on Q.

Using (77) one can calculate an analogous expression for the Riemann curvature tensor on Q

from the usual expression

Rαβγδ =

∂Γαδβ

∂qγ−∂Γαγβ

∂qδ+ ΓαγλΓ

λγβ − ΓαδλΓ

λγβ , (78)

using ∂/∂qγ =∑

i(∂qi/∂qγ) · ∂/∂qi. This gives the curvature written in terms of the original

Euclidean coordinates, the constraints, and the chosen coordinates on Q.

D. d + 1 field theory

The techniques of Sections II A, II B, and II C have natural extensions to field theory.

Given independent field variables ΨA(µ, t), indexed byA = 1, 2, . . . , `, where the independent

variable µ = (µ1, µ2, . . . , µd). The field theoretic version of Hamilton’s principle of (8) is embodied

15

Page 16: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

in the action

S [Ψ] =

∫ t1

t0dt L[Ψ, Ψ] , with L[Ψ, Ψ] =

∫ddµL(Ψ, Ψ, ∂Ψ) , (79)

where we leave the domain of µ and the boundary conditions unspecified, but freely drop surface

terms obtained upon integration by parts. The Lagrangian density L is assumed to depend on the

field components Ψ and ∂Ψ, which is used to indicate all possible partial derivatives with respect

of the components of µ. Hamilton’s principle with (79) gives the Euler-Lagrange equations,

δS [Ψ]δΨA(µ, t)

= 0 ⇒ddt

∂L∂ΨA

+∂

∂µ

∂L∂∂ΨA

−∂L∂ΨA

= 0 , (80)

where the overdot implies differentiation at constant µ. Local holonomic constraints CA(Ψ, ∂Ψ)

are enforced by Lagrange’s method by amending the Lagrangian

Lλ[Ψ, Ψ] =

∫ddµ

(L(Ψ, Ψ, ∂Ψ) + λACA(Ψ, ∂Ψ)

), (81)

with again A = 1, 2, . . . ,M and proceeding as in the finite-dimensional case.

In the Hamiltonian field theoretic setting, we could introduce the conjugate momentum densi-

ties πA,A = 1, 2, . . . , `, with the phase space action

Sλ[Ψ, π] =

∫ t1

t0dt∫

ddµ[πAΨA −H + λADA

], (82)

with Hamiltonian density H and local constraints Da depending on the values of the fields and

their conjugates. Instead of following this route we will jump directly to a generalization of the

field theoretic Dirac bracket formalism that would result.

Consider a Poisson algebra composed of functionals of field variables χA(µ, t) with a Poisson

bracket of the form

{F,G} =

∫ddµ Fχ · J(χ) ·Gχ , (83)

where Fχ is a shorthand for the functional derivative of a functional F with respect to the field χ

[see e.g. 72] and Fχ · J ·Gχ = FχA JABGχB , again with repeated indices summed. Observe the fields

χA(µ, t) need not separate into coordinates and momenta, but if they do the Poisson operator J has

a form akin to that of (23). By a Poisson algebra we mean a Lie algebra realization on functionals,

meaning the Poisson bracket is bilinear, antisymmetric, and satisfies the Jacobi identity and that

there is an associative product of functionals that satisfies the Leibniz law. From the Poisson

bracket the equations of motion are given by χ = {χ,H}, for some Hamiltonian functional H[χ].

16

Page 17: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Dirac’s constraint theory is generally implemented in terms of canonical Poisson brackets [see

e.g. 62, 82, 83], but it is not difficult to show that his procedure also works for noncanonical

Poisson brackets [see e.g. an Appendix of 75].

We impose an even number of local constraints which we write as Da(µ) = constant, a shorthand

for Da[χ(µ)], with the index a = 1, 2, . . . , 2M, bearing in mind that they depend on the fields χ and

their derivatives. As in the finite-dimensional case, the Dirac bracket is obtained from the matrix

D obtained from the bracket of the constraints,

Dab(µ, µ′) = {Da(µ),Db(µ′)} ,

where we note that Dab(µ, µ′) = −Dba(µ′, µ). If D has an inverse, then the Dirac bracket is defined

as follows:

{F,G}∗ = {F,G} −∫

ddµ

∫ddµ′ {F,Da(µ)}D−1

ab (µ, µ′){Db(µ′),G} , (84)

where the coefficients D−1ab (µ, µ′) satisfy∫

ddµ′D−1ab (µ, µ′)Dbc(µ′, µ′′) =

∫d3µ′Dcb(µ, µ′)D−1

ba (µ′, µ′′) = δcaδ(µ − µ

′′) ,

consistent with D−1ab (µ, µ′) = −D−1

ba (µ′, µ).

We note, the procedure is effective only when the coefficients D−1ab (µ, µ′) can be found. If D is

not invertible, then one needs, in general, secondary constraints to determine the Dirac bracket.

1. Field theoretic geodesic flow

In light of Sec. II A 1, any field theory with a Lagrangian density of the form

L =12

ΨA(µ, t) ηAB ΨB(µ, t) , (85)

with the metric ηAB = δAB being the Kronecker delta, subject to time-independent holonomic

constraints can be viewed as geodesic flow on the constraint surface. This is a natural infinite-

dimensional generalization of the idea of Section II A 1.

III. UNCONSTRAINED HAMILTONIAN AND ACTION FOR FLUID

A. Fluid action in Lagrangian variable description

The Lagrangian variable description of a fluid is described in Lagrange’s famous work [69],

while historic and additional material can be found in Morrison [72], Newcomb [76], Serrin [80],

17

Page 18: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Van Kampen & Felderhof [85]. Because the Lagrangian description treats a fluid as a continuum of

particles, it naturally is amenable to the Hamiltonian form. The Lagrangian variable is a coordinate

that gives the position of a fluid element or parcel, as it is sometimes called, at time t. We denote

this variable by q = q(a, t) = (q1, q2, q3), which is measured relative to some cartesian coordinate

system. Here a = (a1, a2, a3) denotes the fluid element label, which is often defined to be the

position of the fluid element at the initial time, a = q(a, 0), but this need not always be the case.

The label a is a continuum analog of the discrete index that labels a generalized coordinate in

a finite degree-of-freedom system. If D is a domain that is fully occupied by the fluid, then at

each fixed time t, q : D → D is assumed to be 1-1 and onto. Not much is really known about

the mathematical properties of this function, but we will assume that it is as smooth as it needs to

be for the operations performed. Also, we will assume we can freely integrate by parts dropping

surface terms and drop reference to D in our integrals.

When discussing the ideal fluid and MHD we will use repeated sum notation with upper and

lower indices even though we are working in cartesian coordinates. And, unlike in Section 2, latin

indices, i, j, k, ` etc. will be summed over 1,2, and 3, the cartesian components, rather than to N.

This is done to avoid further proliferation of indices and we trust confusion will not arise because

of context.

Important quantities are the deformation matrix, ∂qi/∂a j and its Jacobian determinant J :=

det(∂qi/∂a j), which is given by

J =16εk j`ε

imn∂qk

∂ai

∂q j

∂am

∂q`

∂an , (86)

where εi jk = ε i jk is the purely antisymmetric (Levi-Civita) tensor density. We assume a fluid

element is uniquely determined by its label for all time. Thus, J , 0 and we can invert q = q(a, t)

to obtain the label associated with the fluid element at position x at time t, a = q−1(x, t). For

coordinate transformations q = q(a, t) we have

∂qk

∂a j

Aik

J= δi

j , (87)

where Aik is the cofactor of ∂qk/∂ai , which can be written as follows:

Aik =

12εk j`ε

imn ∂q j

∂am

∂q`

∂an . (88)

Using q(a, t) or its inverse q−1(x, t), various quantities can be written either as a function of x

18

Page 19: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

or a. For convenience we list additional formulas below for latter use:

J =13

Ak`

∂q`

∂ak , (89)

A ji =

∂J

∂(∂qi/∂a j), (90)

∂(Aki f )

∂ak = Aki∂ f∂ak , (91)

δJ = Aki∂δqi

∂ak or J = Aki∂qi

∂ak , (92)

δ

(Ak`

J

)∂q`

∂au = −Ak

i

J

∂au δqi or δ

(Ak`

J

)= −

Aki Au

`

J2

∂au δqi , (93)

Au`

∂au

[Ak

i

J

∂z`

∂ak

]= Ak

i∂

∂ak

[Au`

J

∂z`

∂au

], ∀ z` , (94)

which follow from the standard rule for differentiation of determinants and the expression for the

cofactor matrix.

Now we are in a position to recreate and generalize Lagrange’s Lagrangian for the ideal fluid

action principle. On physical grounds we expect our fluid to possess kinetic and internal energies,

and if magnetized, a magnetic energy. The total kinetic energy functional of the fluid is naturally

given by

T [q] :=12

∫d3a ρ0(a) |q|2 , (95)

where ρ0 is the mass density attached to the fluid element labeled by a and q denotes time differ-

entiation of q at fixed label a. Note, in (95) |q|2 = qiqi, where in general qi = gi j qi, but we will

only consider the cartesian metric where gi j = δi j = ηi j.

Fluids are assumed to be in local thermodynamic equilibrium and thus can be described by a

function U(ρ, s), an internal energy per unit mass that depends on the specific volume ρ−1 and

specific entropy s. If a magnetic field B(x, t) were present, then we could add dependence on |B|

as in Morrison [71] to account for pressure anisotropy. The internal energy is written in terms

of the Eulerian density and entropy (see Section III C) since we expect the fluid at each Eulerian

observation point to be in thermal equilibrium. From U we compute the temperature and pressure

according to the usual differentiations, T = ∂U/∂s and p = ρ2∂U/∂ρ. For MHD, the magnetic

energy HB =∫

d3x |B|2/2 in Lagrangian variables would be added. For the ideal fluid, the total

internal energy functional is

V[q] :=∫

d3a ρ0 U (ρ0/J , s0) . (96)

19

Page 20: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Here we have used the fact that a fluid element carries a specfic entropy s = s0(a) and a mass

determined by ρ = ρ0(a)/J . In Section III C we will describe in detail the map from Lagrangian

to Eulerian variables.

Thus, the special case of the action principle of (79) for the ideal fluid has Lagrange’s La-

grangian L[q, q] = T − V . Variation of this action gives the Lagrangian equation of motion for the

fluid

ρ0qi = −A ji∂

∂a j

(ρ2

0

J2

∂U∂ρ

), (97)

with an additional term that describes the J × B force in Lagrangian variables for MHD. See, e.g.,

Morrison [72, 73], Newcomb [76] for details of this calculation and the MHD extension.

B. Hamiltonian formalism in Lagrangian description

Upon defining the momentum density as usual by

πi =δLδqi = ρ0 qi , (98)

we can obtain the Hamiltonian by Legendre transformation, yielding

H[π,q] = T + V =

∫d3a

(|π|2

2ρ0+ ρ0U (ρ0/J , s0)

), (99)

where |π|2 = πiπi = πiηi jπ j. This Hamiltonian with the canonical Poisson bracket,

{F,G} =

∫d3a

(δFδqi

δGδπi−δGδqi

δFδπi

), (100)

yields

qi = {qi,H} = πi/ρ0 , (101)

πi = {πi,H} = −A ji∂

∂a j

(ρ2

0

J2

∂U∂ρ

). (102)

Equations (101) and (102) are equivalent to (97). For MHD a term HB is added to (99) [see 73, 76].

We will give this explicitly in the constraint context in Section IV B 1 after discussing the Lagrange

to Euler map.

C. Hamiltonian formalism in Eulerian description via the Lagrange to Euler map

In order to understand how constraints in terms of the Lagrangian variable description relate

to those in terms of the Eulerian description, in particular ∇ · v = 0, it is necessary to understand

20

Page 21: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

the mapping from Lagrangian to Eulerian variables. Thus, we record here the relationship be-

tween the two unconstrained descriptions, i.e., how the noncanonical Hamiltonian structure of the

compressible Euler’s equations relates to the Hamiltonian structure described in Section III B.

For the ideal fluid, the set of Eulerian variables can be taken to be {v, ρ, s}, where v(x, t) is the

velocity field at the Eulerian observation point, x = (x, y, z) = (x1, x2, x3) at time t and, as as noted

in Section III A, ρ(x, t) is the mass density and s(x, t) is the specific entropy. In order to describe

magnetofluids the magnetic field B(x, t) would be appended to this set. It is most important to

distinguish between the Lagrangian fluid element position and label variables, q and a, and the

Eulerian observation point x, the latter two being independent variables. Confusion exists in the

literature because some authors use the same symbol for the Lagrangian coordinate q and the

Eulerian observation point x.

The Lagrangian and Eulerian descriptions must clearly be related and, indeed, knowing q(a, t)

we can obtain v(x, t). If one were to insert a velocity probe into a fluid at (x, t) then one would

measure the velocity of the fluid element that happened to be at that position at that time. Thus it

is clear that q(a, t) = v(x, t), where recall the overdot means the time derivative at constant a. But,

which fluid element will be at x at time t? Evidently x = q(a, t), which upon inversion yields the

label of that element that will be measured, a = q−1(x, t). Thus, the Eulerian velocity field is given

by

v(x, t) = q(a, t)|a=q−1(x,t) . (103)

Properties can be attached to fluid elements, just as a given mass is identified with a given particle

in mechanics. For a continuum system it is natural to attach a mass density, ρ0(a), to the element

labeled by a. Whence the element of mass in a given volume is given by ρ0d3a and this amount of

mass is preserved by the flow, i.e. ρ(x, t)d3x = ρ0(a)d3a. Because the locus of points of material

surfaces move with the fluid are determined by q, an initial volume element d3a maps into a

volume element d3x at time t according to

d3x = Jd3a , (104)

Thus, using (104) we obtain ρ0 = ρJ as used in Section III A.

Other quantities could be attached to a fluid element; for the ideal fluid, entropy per unit mass,

s(x, t), is such a quantity. The assumption that each fluid element is isentropic then amounts to

s = s0. Similarly, for MHD a magentic field, B0(a), can be attached, and then the frozen flux

assumption yields B · d2x = B0 · d2a. An initial area element d2a maps into an area element d2x at

21

Page 22: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

time t according to

(d2x)i = A ji (d2a) j . (105)

Using (105) we obtain JBi = B j0 ∂qi/∂a j.

Sometimes it is convenient to use another set of Eulerian density variables: {M, ρ, σ,B}, where

σ = ρs is the entropy per unit volume, and M = ρv is the momentum density. These Eulerian

variables can be represented by using the Dirac delta function to ‘pluck out’ the fluid element that

happens to be at the Eulerian observation point x at time t. For example, the mass density ρ(x, t)

is obtained by

ρ(x, t) =

∫d3a ρ0(a) δ (x − q (a, t)) =

ρ0

J

∣∣∣∣∣a=q−1(x,t)

. (106)

The density one observes at x at time t will be the one attached to the fluid element that happens

to be there then, and this fluid element has a label given by solving x = q(a, t). The second

equality of (106) is obtained by using the three-dimensional version of the delta function identity

δ( f (x)) =∑

i δ(x − xi)/| f ′(xi)|, where f (xi) = 0. Similarly, the entropy per unit volume is given by

σ(x, t) =

∫d3aσ0(a) δ (x − q (a, t)) =

σ0

J

∣∣∣∣∣a=q−1(x,t)

, (107)

which is consistent with σ0(a) = ρ0(a)s0(a) and s(x, t) = s0(a)|a=q−1(x,t), where the latter means s

is constant along a Lagrangian orbit. Proceeding, the momentum density, M = (M1,M2,M3), is

related to the Lagrangian canonical momentum (defined in Sec. III B) by

M(x, t) =

∫d3aπ(a, t) δ (x − q(a, t)) =

π(a, t)J

∣∣∣∣∣a=q−1(x,t)

, (108)

where for the ideal fluid and MHD, π(a, t) = (π1, π2, π3) = ρ0q. Lastly,

Bi(x, t) =

∫d3a

∂qi(a, t)∂a j B j

0(a) δ (x − q(a, t)) =∂qi(a, t)∂a j

B j0(a)J

∣∣∣∣∣∣∣a=q−1(x,t)

, (109)

for the components of the magnetic field. It may be unfamiliar to view the magnetic field as

density, but in MHD it obeys a conservation law. Geometrically, however, it is naturally a vector

density associated with a differential 2-form.

To obtain the noncanonical Eulerian Poisson bracket we consider functionals F[q, π] that are

restricted so as to obtain their dependence on q and π only through the Eulerian variables. Upon

setting F[q,π] = F[v, ρ, σ], equating variations of both sides,

δF =

∫d3a

[δFδq· δq +

δFδπ· δπ

]=

∫d3x

[δFδρδρ +

δFδσ

δσ +δFδM· δM

]= δF , (110)

22

Page 23: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

varying the expressions (106), (107), and (108), substituting the result into (110), and equating the

independent coefficients of δq and δπ, we obtain

δFδq

=

∫d3x

[ρ0 ∇

δFδρ

+ σ0 ∇δFδσ

+ πi∇δFδMi

]δ(x − q) , (111)

δFδπ

=

∫d3x

δFδM

δ(x − q) . (112)

(See [72] and [74] for details.) Upon substitution of (111) and (112), expressions of the functional

chain rule that relate Lagrangian functional derivatives to the Eulerian functional derivates, into

(100) yields the following bracket expressed entirely in terms of the Eulerian fields {M, ρ, σ}:

{F,G} = −

∫d3x

[Mi

(δFδM j

∂x j

δGδMi−δGδM j

∂x j

δFδMi

)+ ρ

(δFδM· ∇

δGδρ−δGδM· ∇

δFδρ

)+ σ

(δFδM· ∇

δGδσ−δGδM· ∇

δFδσ

)]. (113)

In (113) we have dropped the overbars on the Eulerian functional derivatives. The bracket for

MHD is the above with the addition of the following term, which is obtained by adding a B

contribution to (110):

{F,G}B = −

∫d3x

[B ·

(δFδM· ∇

δGδB−δGδM· ∇

δFδB

)+ B ·

(∇

(δFδM

)·δGδB− ∇

(δGδM

)·δFδB

) ], (114)

where dyadic notation is used; for example, B · [∇(D) · C] =∑

i, j BiC j∂D j/∂xi, for vectors B,D,

and C. Alternatively, the bracket in terms of {v, ρ, s,B} is obtained using chain rule expressions,

e.g.,δFδρ

∣∣∣∣∣∣v,s

=δFδρ

∣∣∣∣∣∣M,s

+Mρ·δFδM

ρ

δFδσ

, (115)

yielding

{F,G} = −

∫d3x

[(δFδρ∇ ·

δGδv−δGδρ∇ ·

δFδv

)+

(∇ × vρ·δGδv×δFδv

)+∇sρ·

(δFδsδGδv−δGδs

δFδv

)], (116)

and

{F,G}B = −

∫d3x

[B ·

(1ρ

δFδv· ∇

δGδB−

δGδv· ∇

δFδB

)+ B ·

(∇

(1ρ

δFδv

)·δGδB− ∇

(1ρ

δGδv

)·δFδB

) ]. (117)

23

Page 24: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

The bracket of (116) plus that of (117) with the Hamiltonian

H[ρ, s, v,B] =

∫d3x

(12ρ|v|2 + ρU(ρ, s) +

12|B|2

)(118)

gives the Eulerian version of MHD in Hamiltonian form, ∂v/∂t = {v,H}, etc., and similarly using

(113) plus (114) with the Hamiltonian expressed in terms of (M, ρ, σ,B). Ideal fluid follows upon

neglecting the B terms.

D. Constants of motion: Eulerian vs. Lagrangian

In oder to compare the imposition of constraints in the Lagrangian and Eulerian descriptions, it

is necessary to compare Lagrangian and Eulerian conservations laws. This is because constraints,

when enforced, are conserved quantities. The comparison is not trivial because time independent

quantities in the Eulerian description can be time dependent in the Lagrangian description.

Consider a Lagrangian function f (a, t), typical of the Lagrangian variable description, and the

relation x = q(a, t), which relates an Eulerian observation point x to a corresponding fluid element

trajectory value. The function f can be written in either picture by composition, as follows:

f (a, t) = f (x, t) = f (q(a, t), t) , (119)

where we will use a tilde to indicated the Eulerian form of a Lagrangian function. Application of

the chain rule gives

Aik

J

∂ f∂ai

∣∣∣∣∣∣a=q−1(x,t)

=∂ f∂xk and

Ak`

J

∂ak

(π`

J

)∣∣∣∣∣∣a=q−1(x,t)

= ∇ · v , (120)

with the second equality of (120) being a special case of the first. Similarly,

f∣∣∣a=q−1(x,t) =

∂ f∂t

+ qi(a, t)∂ f∂xi

∣∣∣∣∣∣a=q−1(x,t)

=∂ f∂t

+ v · ∇ f (x, t) , (121)

where recall an overdot denotes the time derivative at constant a, ∂/∂t denotes the time derivative

at constant x, and ∇ is the Eulerian gradient with components ∂/∂xi as used in (120). Because

the Jacobian determinant J is composed of derivatives of q, we have J(a, t)|a=q−1(x,t) = J(x, t),

whence one obtains a formula due to Euler [see e.g. 80],

∂J

∂t+ v · ∇J = J ∇ · v . (122)

24

Page 25: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Now, consider a conservation law in the Lagrangian variable description,

DL +∂ΓiDL

∂ai = 0 , (123)

where the densityDL(a, t) has the associated flux ΓDL . Then, the associated conserved quantity is

IDL =

∫d3aDL , (124)

which satisfies dIDL/dt = 0 provided surface terms vanish. Similarly, an Eulerian conservation

law with densityDE and flux ΓDE is

∂DE

∂t+∂ΓiDE

∂xi = 0 (125)

and the following is similarly constant in time:

IDE =

∫d3xDE . (126)

The relationship between the two conservation laws (123) and (125) can be obtained by defining

DL = JDE , ΓiDL

= Aik ΓkDE, and ΓDE = ΓDE + vDE , (127)

and their equivalence follows from (92), (121), and (122). Given a Lagrangian conservation law,

one can use (127) to obtain a corresponding Eulerian conservation law. The densityDE is obtained

from the first equation of (127), a piece of the Eulerian flux ΓDE from the second, which then can

be substitued into the third equation of (127) to obtain the complete Eulerian flux ΓDE . An Eulerian

conservation law is most useful when one can write DE and ΓDE entirely in terms of the Eulerian

variables of the fluid.

The simplest case occurs when DL only depends on a, in which case the corresponding flux

is zero and ∂DL/∂t = 0 and dIDL/dt = 0 follow directly because (124) has no time dependence

whatsoever. Any attribute attached to a fluid element only depends on the label a and this has a

trivial conservation law of this form. However, such trivial Lagrangian conservation laws yield

nontrivial Eulerian conservation laws. Observe, even thought ΓDE ≡ 0 by (127), ΓDE = vDE , 0.

Consider the case of the entropy whereDL = s0(a), whence s(x, t) = s0(a(x, t)) and by (121),

∂s∂t

+ v · ∇s = 0 , (128)

with the quantity s = s0/J being according to (127) the Eulerian conserved density, as can be

verified using (122). But, as it stands, this density cannot be written in terms of Eulerian fluid

25

Page 26: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

variables. However, σ0 = ρ0s0 is also a trivial Lagrangian conserved density and according to

(127) we have the Eulerian density ρ0s0/J = ρs = σ that satisfies

∂σ

∂t+ ∇ · (vσ) = 0 . (129)

Thus, it follows that any advected scalar has an associated conserved quantity obtained by multi-

plication by ρ.

As another example, consider the quantity Bi0∂q j/∂ai. This quantity is the limit displacement

between two nearby fluid elements, i.e., q(a, t) − q(a + δa, t) along the initial magnetic field as

δa→ 0. Evidently,˙(

Bi0∂q j

∂ai

)= Bi

0∂q j

∂ai =∂

∂ai

(Bi

0q j), (130)

where the second equality follows if the initial magnetic field is divergence free. This is of course

another trivial conservation law, for the time derivative of a density that is a divergence will always

be a divergence. However, let us see what this becomes in the Eulerian description. According

to (127) the corresponding Eulerian density is DE = DL/J ; so, the density associated with this

trivial conservation law (130) is

B j(x, t) =Bi

0

J

∂q j

∂ai

∣∣∣∣∣∣a=q−1(x,t)

. (131)

which as we saw in Section III C is the expression one gets for the MHD magnetic field because

of flux conservation. That the divergence-free magnetic field satisfies a conservation law is clear

from∂B∂t

= −v · ∇B + B · ∇v − B∇ · v = ∇ ·↔

T , (132)

where the tensor↔

T of the last equality is

T = B ⊗ v − v ⊗ B . (133)

Thus we have another instance where a trivial Lagrangian conservation law leads to a nontrivial

Eulerian one.

Although Bi0∂q j/∂ai does not map into an expression entirely in terms of our set of Eulerian

variables, we can divide it by ρ0, a quantity that only depends on the label a, and obtain

Bi0

ρ0

∂q j

∂ai

∣∣∣∣∣∣a=q−1(x,t)

=B j

ρ. (134)

26

Page 27: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Eulerianizing the counterpart of (130) for this expression gives

∂t

(Bρ

)+ v · ∇

(Bρ

)=

Bρ· ∇v , (135)

which is not an Eulerian conservation law. This is to be expected because, unlike what we did to get

(132), we have Eulerianized without using (127). In light of its relationship to q(a, t)−q(a +δa, t),

the quantity B/ρ has been described as a measure of the distance of points on a magnetic field line

[see e.g. 68]. This was predated by analogous arguments for vorticity [see e.g. 80].

IV. CONSTRAINT THEORIES FOR THE INCOMPRESSIBLE IDEAL FLUID

A. The incompressible fluid in Lagrangian variables

In order to enforce incompressibility, Lagrange added to his Lagrangian the constraint J = 1

with the Lagrange multiplier λ(a, t),

Lλ[q, q] = T [q] + λJ , (136)

with T given (95). Here we have dropped V because incompressible fluids contain no internal

energy. Upon insertion of (136) into the action of Hamilton’s principle it is discovered that λ

corresponds to the pressure. The essence of this procedure was known to Lagrange. (See Serrin

[80] for historical details and Sommerfeld [81] for an elementary exposition.) This procedure

yields

ρ0qi = −Aij∂λ

∂a j , (137)

where use has been made of (92). The Eulerian form of (137) is clearly ρ(∂v/∂t + v · ∇v) = −∇λ,

whence it is clear that λ is the pressure. Although Lagrange knew the Lagrange multiplier was the

pressure, he could only solve for it in special cases. The general procedure of Section I A was not

available because Green’s function techniques and the theory of elliptic equations were not at his

disposal.

1. Lagrangian volume preserving geodesic flow

If the constraint is dropped from (136), we obtain free particle motion for an infinite-dimensional

system, the ideal fluid case of (85) of Section II D, which is analogous to the finite-dimensional

27

Page 28: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

case of Section II A 1. Because the constraint J = 1 only depends on the derivatives of q, it is a

configuration space constraint; thus, it is an holonomic constraint. As is well-known and reviewed

in Section II A 1, free particle motion with holonomic constaints is geodesic flow. Thus, following

Lagrange, it is immediate that the ideal incompressible fluid is an infinite-dimensional version of

geodesic flow.

Lagrange’s calculation was placed in a geometric/group theoretic setting in [50] (see also Ap-

pendix 2 of [51] and [52]). Given that the transformation a ↔ q, at any time, is assumed to be

a smooth invertible coordinate change, it is a Lie group, one referred to as the diffeomorphism

group. With the additional assumption that these transformations are volume preserving, La-

grange’s constraint J = 1, the transformations form a subgroup, the group of volume preserving

diffeomorphisms. Thus, Lagrange’s work can be viewed as geodesic flow on the group of volume

preserving diffeomorphisms.

Although Arnold’s assumptions of smoothness etc. are mathematically dramatic, his descrip-

tion of Lagrange’s calculations in these terms has spawned a considerable body of research. Asso-

ciated with a geodesic flow is a metric, and whence one can calculate a curvature. In his original

work, Arnold added the novel calculation of the curvature in the mathematically more forgiving

case of two-dimensional flow with periodic boundary conditions.

B. Lagrangian-Dirac constraint theory

More recently there have been several works [55, 75, 84], following Nguyen & Turski [77, 78],

that treat the enforcement of the incompressibility constraint of hydrodynamics by Dirac’s method

of constraints [62]. In these works the compressibility constraint was enforced in the Eulerian

variable description of the fluid using the noncanonical Poisson bracket of Section III C as the base

bracket of a generalization of Dirac’s constraint theory. We will return to this approach in Section

III C where we revisit and extend Dirac’s constraint results for the fluid in the Eulerian variable

description. Here, apparently for the first time, we consider the incompressibility constraint in the

Lagrangian variable description, where the canonical Poisson bracket of (100) is the base for the

construction of a Dirac bracket.

We adapt (84) for the fluid case at hand with the supposition of only two local constraints,

which we write as

Da(a′) =

∫d3a Da(a) δ(a − a′) , (138)

28

Page 29: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

where a = 1, 2 and Da(a) is a shorthand for a function of q(a, t) and π(a, t) and their derivatives

with respect to a. Then the matrix D is a 2 × 2 matrix with the components

Dab(a, a′) = {Da(a),Db(a′)} ,

using the canonical bracket of (100). To construct the Dirac bracket

{F,G}∗ = {F,G} −∫

d3a∫

d3a′ {F,Da(a)}D−1ab (a, a′){Db(a′),G}, (139)

we require the inverse, which satisfies∫d3a Dac(a′, a)D−1

cb (a, a′′) = δab δ(a

′ − a′′) . (140)

Rather than continuing with the general case, which is unwieldy, we proceed to the special

case for the incompressible fluid, an infinite-dimensional version of the holonomic constraints

discussed in Section II C 1.

1. Lagrangian-Dirac incompressibility holonomic constraint

Evidently we will want our holonomic incompressibility constraint to be J . However, it is

convenient to express this by choosing

D1 = ln(J

ρ0

). (141)

This amounts to the same constraint as J = 1 with the value D1 = − ln(ρ0). The scaling of J

in (141) by ρ0(a) is immaterial because it is a time-independent quantity. To obtain the second

constraint we follow suit and set

D2 = D1 =Ak`

J

∂ak

(π`

ρ0

)= η` j Ak

`

J

∂ak

(π j

ρ0

), (142)

where recall we assume η` j = δ` j and π j is given by (98). That the constraint D2 is the time

derivative of D1 requires the definition of π j of (98) that uses the Hamiltonian∫

d3a |π|2/(2ρ0).

Observe, that constraints D1 and D2 are local constraints in that they are enforced pointwise [see

e.g. 65], i.e., they are enforced on each fluid element labeled by a. Equation (141) corresponds in

the Eulerian picture to − ln(ρ), while the second constraint of (142), the Lagrangian time derivative

of the first constraint, corresponds in the Eulerian picture to ∇ · v, which can be easily verified

using the second equation of (120). Note, the particular values of these constraints of interest are,

29

Page 30: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

of course, J = 1 and ∇ · v = 0, but the dynamics the Dirac bracket generates will preserve any

values of these constraints. For example, we could set J = f (a) where the arbitrary function f is

less than unity for some a and greater for others, corresponding to regions of fluid elements that

experience contraction and expansion. Also note, because we have used π with the up index in

(142); thus as seen in the second equality it depends on the metric. This was done to make it have

the Eulerian form ∇ · v.

For the constraints (141) and (142), D only depends on two quantities because D1 does not

depend on π, i.e. {D1,D1} = 0 and {D1,D2} = −{D2,D1}. Thus the inverse has the form

D−1 =

D−111 D

−112

D−121 0

, (143)

giving rise to the conditions

D−112 · D

21 = I = D−121 · D

12 and D−111 · D

12 + D−112 · D

22 = 0 , (144)

where I is the identity. Thus, the inverse is easily tractable if the inverse of D12 exists; whence,

D−111 = −D−1

12 · D22 · D−1

21 . (145)

In the above the symbol ‘ · ’ is used to denote the product with the sum in infinite dimensions,

i.e., integration over d3a as in (140). Equation (145) can be rewritten in an abbreviated form with

implied integrals on repeated arguments as

D−111 (a′, a′′) = D−1

21 (a′, a) · D22(a, a) · D−121 (a, a′′) . (146)

In order to obtain D and its inverse, we need the functional derivatives of D1 and D2. These are

obtained directly by writing these local constraints as in (140), yielding

δD1(a′)δqi(a)

= −Aki∂

∂ak

δ(a − a′)J

, (147)

δD1(a′)δπi(a)

= 0 , (148)

where use has been made of (92), and

δD2(a′)δqi(a)

=∂

∂au

(Ak

i Au`

J2

∂ak

(π`

ρ0

)δ(a − a′)

), (149)

δD2 (a′)δπi(a)

= −ηi j

ρ0

∂am

(Amj

Jδ(a − a′

)), (150)

30

Page 31: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

where use has been made of (93) and recalling we have (91) at our disposal.

Let us now insert (147), (148), (149), and (150) into the canonical Poisson bracket (100), to

obtain

D12(a, a′) = {D1(a),D2(a′)}

= −A`

i

J

∂a`

(ηi j

ρ0Ak

j∂

∂ak

(δ(a − a′)J

)), (151)

which corresponds to the symmetric matrix S of (34) and (49) and

D22(a, a′) = {D2(a),D2(a′)}

=Ak

i Au`

J2

∂ak

(π`

ρ0

)∂

∂au

[ηi j

ρ0

∂am

(Amj

Jδ(a − a′

))]−

Ami

J

∂am

ηi j

ρ0

∂au

AkjA

u`

J2

∂ak

(π`

ρ0

)δ(a − a′)

, (152)

which corresponds to the antisymmetric matrix A of (35). Observe the symmetries corresponding

to the matrices S and A, respectively, are here∫d3a′D12(a, a′) φ(a′) =

∫d3a′D12(a′, a) φ(a′) ,∫

d3a′D22(a, a′) φ(a′) = −

∫d3a′D22(a′, a) φ(a′) ,

for all functions φ. The first follows from integration by parts, while the second is obvious from

its definition.

Using (140), the first condition of (144) is∫d3a′′ D12(a′, a′′)D−1

21 (a′′, a) = δ(a′ − a) , (153)

which upon substitution of (151) and integration gives

−A`

i

J

∂a`

[ηi j

ρ0Ak

j∂

∂ak

(D−1

21 (a, a′′)J

)]= δ(a − a′′) . (154)

We introduce the formally self-adjoint operator (cf. (91))

∆ρ0 f =A`

i

J

∂a`

[ηi j

ρ0Ak

j∂

∂ak

(fJ

)], (155)

i.e., an operator that satisfies∫d3a f (a) ∆ρ0g(a) =

∫d3a g(a) ∆ρ0 f (a) , (156)

31

Page 32: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

a property inherited by its inverse ∆−1ρ0

. Thus we can rewrite equation (154) as

D−121 (a, a′′) = −G0

(a, a′′

)= −∆−1

ρ0δ(a − a′′) , (157)

where G0 represents the Green function associated with (154).

In order to obtain D−121 , we find it convenient to transform (157) to Eulerian variables. Using

x = q(a, t) we findD−1

21 (a, a′)J

= −G(x, x′) = −G(q(a),q(a′)) , (158)

where G satisfies

∇ ·

(1ρ∇G

)= −∆ρ

D−121 (a, a′)J

= Jδ(x − x′). (159)

Here use has been made of identities (91) and (120). As noted in Section I A, under physically

reasonable conditions, the operator

∆ρ f = ∆ρ0 (J f ) = ∇ ·

(1ρ∇ f

)(160)

has an inverse. Thus we write

D−121 (a, a′) = −J∆−1

ρ

(J δ

(q(a, t) − q(a′, t)

)). (161)

Now, using D−121 = −D−1

12 , the element D−111 follows directly from (145).

For convenience we write the Dirac bracket of (139) as follows:

{F,G}∗ = {F,G} − [F,G]D , (162)

where

[F,G]D :=2∑

a,b=1

[F,G]Dab =

∫d3a

∫d3a′ {F,Da (a)}D−1

ab (a, a′){Db (

a′),G

}. (163)

Because D−122 = 0 and [F,G]D

12 = −[G, F]D21, we only need to calculate [F,G]D

11 and [F,G]D21.

As above, we substitute (147), (148), (149), and (150) into the bracket (100) and obtain{F,D1 (a)

}= −

Aki

J

∂ak

(δFδπi

), (164)

{F,D2 (a)

}=

Ak`

J

∂ak

(ηi`

ρ0

δFδqi

)+

Aki

J

Au`

J

∂ak

(π`

ρ0

)∂

∂au

(δFδπi

). (165)

Then, exploiting the antisymmetry of the Poisson bracket, it is straightforward to calculate analo-

gous expressions for the terms{D1,2,G

}.

32

Page 33: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

We first analyze the operator

[F,G]D11 =

∫d3a

∫d3a′

∫d3a

∫d3a

{F,D1(a)

}D−1

21 (a, a)D22(a, a)D−121 (a, a′)

{D1(a′) ,G}

, (166)

where we used the second condition of (144) to replace D−111 . Upon inserting (161) and (164), this

equation can be rewritten as

[F,G]D11 = −

∫d3a

∫d3a′

∫d3a

∫d3aAh

j

J

∂ah

(δFδπ j

)∆−1ρ0δ(a − a)

a=a

D22(a, a)[As

r

J

∂as

(δGδπr

)∆−1ρ0δ(a − a)

]a=a′

, (167)

where the subscripts on the right delimiters indicate that a is to be replaced after the derivative

operations, including those that occur in J and A ji .

Integrating this expression by parts with respect to a and a′ yields

[F,G]D11 = −

∫d3a

∫d3a

∆−1ρ0

Ahj

J

∂ah

(δFδπ j

)a=a

D22(a, a)[∆−1ρ0

(As

r

J

∂as

(δGδπr

))]a=a

, (168)

and then substituting (152) gives

[F,G]D11 = −

∫d3a

∫d3a

∆−1ρ0

Ahj

J

∂ah

(δFδπ j

)a=a

{Ak

i Au`

J2

∂ak

(π`

ρ0

)∂

∂au

[ηin

ρ0

∂am

(Am

n

Jδ (a − a)

)]−

Ami

J

∂am

[ηin

ρ0

∂au

(Ak

n

J

Au`

J

∂ak

(π`

ρ0

)δ(a − a)

)]}a=a

[∆−1ρ0

(As

r

J

∂as

(δGδπr

))]a=a

. (169)

Then, by means of integrations by parts we can remove the derivatives from the term δ(a − a) and

perform the integral. After relabeling the integration variable as a to simplify the notation, (169)

becomes

[F,G]D11 =

∫d3a

{ρ0 η

ui Aku

J

∂ak

(π`

ρ0

) (Pρ0⊥

δFδπ

∣∣∣∣∣`

Pρ0⊥

δGδπ

∣∣∣∣∣i− Pρ0⊥

δFδπ

∣∣∣∣∣iPρ0⊥

δGδπ

∣∣∣∣∣`

)

+ ηniAu`

∂au

[Ak

n

J

∂ak

(π`

ρ0

)] Pρ0⊥

δGδπ

∣∣∣∣∣i

∆−1ρ0

J

Ahj

J

∂ah

(δFδπ j

)− Pρ0⊥

δFδπ

∣∣∣∣∣i

∆−1ρ0

J

Ahj

J

∂ah

(δGδπ j

)}, (170)

where we introduced the projection operator

(Pρ0⊥

)i

jz j =

ηi`

ρ0Au`

∂au

∆−1ρ0

J

Ahj

J

∂ah z j

=: Pρ0⊥z∣∣∣i , (171)

33

Page 34: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

where in the last equality we defined a shorthand for convenience; similarly we use

Pρ0⊥z∣∣∣`

:=1ρ0

Au`

∂au

∆−1ρ0

J

Ahj

J

∂ah z j

. (172)

It is straightforward to prove that Pρ0⊥ represents a projection, i.e. Pρ0⊥

(Pρ0⊥z

)= Pρ0⊥z for each

z, which in terms of indices would have an ith component given by (Pρ0⊥)ij(Pρ0⊥) j

k zk = (Pρ0⊥)ik zk.

Also, Pρ0⊥ is formally self-adjoint with respect to the following weighted inner product:∫d3a ρ0 wi (Pρ0⊥)i

j z j =

∫d3a ρ0 zi (Pρ0⊥)i

j w j . (173)

The projection operator complementary to Pρ0⊥ is given by

Pρ0 = I − Pρ0⊥ , (174)

where I is the identity.

Now let us return to our evaluation of [F,G]D and analyze the contribution

[F,G]D21 =

∫d3a

∫d3a′

{F,D2 (a)

}D−1

21 (a, a′){D1 (

a′),G

}. (175)

Using (161), (164), and (165), this equation can be rewritten as

[F,G]D21 = −

∫d3a

∫d3a′

[Ak

i

J

∂ak

(ηin

ρ0

δFδqn

)+

Aki

J

Au`

J

∂ak

(π`

ρ0

)∂

∂au

(δFδπi

)]× ∆−1

ρ0δ(a − a′)

Ahj

J

∂ah

(δGδπ j

)a=a′

(176)

and, integrating by parts to simplify the δ(a − a′) term, results in

[F,G]D21 =

∫d3a

{δFδqi Pρ0⊥

δGδπ

∣∣∣∣∣i + ρ0Ak

i

J

∂ak

(π`

ρ0

)δFδπiPρ0⊥

δGδπ

∣∣∣∣∣`

+ Au`

∂au

[Ak

i

J

∂ak

(π`

ρ0

)]δFδπi

∆−1ρ0

J

Ahj

J

∂ah

(δGδπ j

)}. (177)

We can now combine the operators [F,G]D11, [F,G]D

21, and [F,G]D12 = −[G, F]D

21, given by (170)

and (177), to calculate the Dirac bracket (162). First, we rewrite (163) as

[F,G]D =

∫d3a

{δFδqi Pρ0⊥

δGδπ

∣∣∣∣∣i − δGδqi Pρ0⊥

δFδπ

∣∣∣∣∣i+ ρ0

Aki

J

∂ak

(π`

ρ0

) (Pρ0

δFδπ

∣∣∣∣∣i Pρ0⊥

δGδπ

∣∣∣∣∣`

− Pρ0

δGδπ

∣∣∣∣∣i Pρ0⊥

δFδπ

∣∣∣∣∣`

)+ Au

`

∂au

[Ak

i

J

∂ak

(π`

ρ0

)] Pρ0

δFδπ

∣∣∣∣∣i ∆−1ρ0

J

Ahj

J

∂ah

(δGδπ j

)− Pρ0

δGδπ

∣∣∣∣∣i ∆−1ρ0

J

Ahj

J

∂ah

(δFδπ j

)}. (178)

34

Page 35: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Using the identity of (94) with z` set to π`/ρ0,

Au`

∂au

[Ak

i

J

∂ak

(π`

ρ0

)]= Ak

i∂

∂ak

[Au`

J

∂au

(π`

ρ0

)], (179)

and integrating by parts, (178) becomes

[F,G]D =

∫d3a

{δFδqi Pρ0⊥

δGδπ

∣∣∣∣∣i − δGδqi Pρ0⊥

δFδπ

∣∣∣∣∣i+ ρ0

Aki

J

∂ak

(π`

ρ0

) (Pρ0

δFδπ

∣∣∣∣∣i Pρ0⊥

δGδπ

∣∣∣∣∣`

− Pρ0

δGδπ

∣∣∣∣∣i Pρ0⊥

δFδπ

∣∣∣∣∣`

)− ρ0

Au`

J

∂au

(π`

ρ0

) (Pρ0

δFδπ

∣∣∣∣∣i Pρ0⊥

δGδπ

∣∣∣∣∣i− Pρ0

δGδπ

∣∣∣∣∣i Pρ0⊥

δFδπ

∣∣∣∣∣i

)}, (180)

where we used

Aki∂

∂akPρ0z∣∣∣i = Ak

i∂

∂ak

(Pρ0

)i

jz j = 0 , for all z , (181)

which follows from the definitions (171), viz.

Aki

J

∂ak

(Pρ0⊥

)i

jz j =

Aki

J

∂z j

∂ak , (182)

and (174). Also, upon inserting Pρ0⊥ = I − Pρ0 in the last line of (180), symmetry implies we can

drop the Pρ0⊥. Finally, upon substituting (180) into (162), we obtain

{F,G}∗ =

∫d3a

{δFδqi Pρ0

δGδπ

∣∣∣∣∣i − δGδqi Pρ0

δFδπ

∣∣∣∣∣i− ρ0

Aki

J

∂ak

(π`

ρ0

) (Pρ0

δFδπ

∣∣∣∣∣i Pρ0⊥

δGδπ

∣∣∣∣∣`

− Pρ0

δGδπ

∣∣∣∣∣i Pρ0⊥

δFδπ

∣∣∣∣∣`

)+ ρ0

Ak`

J

∂ak

(π`

ρ0

) (Pρ0

δFδπ

∣∣∣∣∣i δGδπi − Pρ0

δGδπ

∣∣∣∣∣i δFδπi

)}. (183)

Once more inserting Pρ0⊥ = I − Pρ0 , rearranging, and reindexing gives

{F,G}∗ = −

∫d3a ρ0

{1ρ0

δGδqi Pρ0

δFδπ

∣∣∣∣∣i − 1ρ0

δFδqi Pρ0

δGδπ

∣∣∣∣∣i +Amn Pρ0⊥

δFδπ

∣∣∣∣∣m Pρ0⊥

δGδπ

∣∣∣∣∣n+ Tmn

(δFδπmPρ0⊥

δGδπ

∣∣∣∣∣n − δGδπmPρ0⊥

δFδπ

∣∣∣∣∣n)}, (184)

where

Anm := η`mD`n − η`nD`

m and Tmn := η`nD`m + ηmnD2 , (185)

35

Page 36: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

with

D`m =

Akm

J

∂ak

(π`

ρ0

). (186)

Note the trace D`` = D2, which we will eventually set to zero. Equation (184) gives the Dirac

bracket for the incompressibility holonomic constraint. This bracket with the Hamiltonian

H =

∫d3a|π|2

2ρ0=

∫d3a ηmnπmπn

2ρ0, (187)

produces dynamics that fixesJ and thus enforces incompressibility provided the constraint D2 = 0

is used as an initial condition. For MHD we add to H the following:

HB =

∫d3a ηmn

B j0Bk

0

2J∂qm

∂a j

∂qn

∂ak . (188)

We note, any Hamiltonian that is consistent with (142) can be used to define a constrained flow.

Proceeding to the equations of motion, we first calculate qi,

qi = {qi,H}∗ =(Pρ0

)i

j

δHδπ j

=δHδπi−ηi`

ρ0Au`

∂au

∆−1ρ0

J

Ahj

J

∂ah

δHδπ j

=πi

ρ0−ηi`

ρ0Au`

∂au

∆−1ρ0

J

Ahj

J

∂ah

π j

ρ0

. (189)

The equation for πi is more involved. Using the adjoint property of (173), which is valid for both

Pρ0⊥ and Pρ0 , we obtain

πi = {πi,H}∗ = −ρ0

(Pρ0

) j

i

1ρ0

δHδq j − ρ0

(Pρ0⊥

)m

i

(Amn

(Pρ0⊥

)n

k

δHδπk

)+ ρ0

(Pρ0⊥

)n

i

(Tmn

δHδπm

)− ρ0 Tin

(Pρ0⊥

)n

k

δHδπk

= −ρ0

(Pρ0⊥

)m

i

(Amn

(Pρ0⊥

)n

k

πk

ρ0

)+ ρ0

(Pρ0⊥

)n

i

(Tmn

πm

ρ0

)− ρ0 Tin

(Pρ0⊥

)n

k

πk

ρ0, (190)

which upon substitution of the definitions of Pρ0 , Amn, and Tmn of (171) and (185) yields a com-

plicated nonlinear equation.

Equations (189) and (190) are infinite-dimensional versions of the finite-dimensional systems

of (40) and (41) considered in Section II C 1. There, equations (40) and (41) were reduced to (44)

and (45) upon enforcing the holomomic constraint by requiring that initially D2 = 0. Similarly

we can enforce the vanishing of D2 of (142), which is compatible with the Hamiltonian (187).

Instead of addressing this evaluation now, we find the meaning of various terms is much more

transparent when written in terms of Eulerian variables, which we do in Section IV C. We then

36

Page 37: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

return to these Lagrangian equations in Section IV D and make comparisons. Nevertheless, the

solution of equations (189) and (190), q(a, t), with the initial conditions D1 = − ln ρ0 and D2 = 0,

is a volume preserving transformation at any time t.

C. Eulerian-Dirac constraint theory

Because we chose the form of constraints D1,2 of (141) and (142) to be Eulerianizable, it follows

that we can transform easily the results of Section IV B 1 into Eulerian form. This we do in Section

IV C 1. Alternatively, we can proceed as in Chandre et al. [55], Morrison et al. [75], Nguyen &

Turski [77, 78], Tassi et al. [84], starting from the Eulerian noncanonical theory of Section III C

and directly construct a Dirac bracket with Eulerian constraints. This is a valid procedure because

Dirac’s construction works for noncanonical Poisson brackets, as shown, e.g., in Morrison et al.

[75], but it does not readily allow for advected density. This direct method with uniform density is

reviewed in Section IV C 2, where it is contrasted with the results of Section IV C 1.

1. Lagrangian-Dirac constraint theory in the Eulerian picture

In a manner similar to that used to obtain (111) and (112), we find the functional derivatives

transform asδFδπi

=1ρ

δFδ3i,

1ρ0

δFδqi =

∂xi

δFδρ−

δFδs

∂s∂xi −

δFδ3`

∂3`∂xi , (191)

where the expressions on the left of each equality are clearly Lagrangian variable quantities, while

on the right they are Eulerian quantities represented in terms of Lagrangian variables. Substituting

these expressions into (84) and dropping the bar on F and G gives the following bracket in terms

of the Eulerian variables:

{F,G}∗ = −

∫d3x

{ (PρδFδv

∣∣∣∣∣i ∂

∂xi

δGδρ− Pρ

δGδv

∣∣∣∣∣i ∂

∂xi

δFδρ

)+

∂s∂xi

(δFδsPρδGδv

∣∣∣∣∣i − δGδsPρδFδv

∣∣∣∣∣i)+

∂3`

∂xi

(PρδFδv

∣∣∣∣∣`

PρδGδv

∣∣∣∣∣i − Pρ δGδv∣∣∣∣∣`

PρδFδv

∣∣∣∣∣i)+

∂3`

∂x`

(PρδFδv

∣∣∣∣∣i δGδ3i − Pρ δGδv∣∣∣∣∣i δFδ3i

)}, (192)

37

Page 38: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

where we used the relations (104) and (120) and we introduced the Eulerian projection operator

PρδFδv

∣∣∣∣∣i = (Pρ)ijδFδ3 j

= ρPρ0

δFδπ

∣∣∣∣∣i and PρδFδv

∣∣∣∣∣i= ηi j Pρ

δFδv

∣∣∣∣∣ j

, (193)

with (Pρ

)i

jz j = δi

j − ηik ∂

∂xk

[∆−1ρ

∂x j

(z j

ρ

)], (194)

which is easily seen to satisfy (Pρ)ij(Pρ)

jk = (Pρ)i

k. Observe, like its Lagrangian counterpart, Pρ is

formally self-adjoint; however, this time we found it convenient to define the projection in such a

way that the self-adjointness is with respect to a different weighted inner product, viz.∫d3xρ

wi (Pρ)ij z j =

∫d3xρ

zi (Pρ)ij w j . (195)

In terms of usual cartesian vector notation

PρδGδv

=δGδv− ∇∆−1

ρ ∇ ·

(1ρ

δGδv

). (196)

Upon writing Pρ = I − Pρ⊥ and decomposing an arbitrary vector field as

z = −∇Φ + ρ∇ × A,

this projection operator yields the component Pρz = ρ∇ × A. Hence, ∇ρ × A = 0 is sufficient for

the operator to be a projection into the space of incompressible vector fields. For convenience we

introduce the associated projector

Pρv := v −1ρ∇∆−1

ρ ∇ · v =1ρPρ(ρv) , (197)

which has the desirable property

∇ · (Pρv) = 0 ∀ v compared to ∇ ·

(1ρPρw

)= 0 ∀ w . (198)

Upon writing Pρ = I − Pρ⊥ and decomposing an arbitrary vector field v as

v = −1ρ∇Φ + ∇ × A,

this projection operator yields the component Pρv = ∇ × A, while Pρ⊥v = ∇Φ/ρ. Note, Pρ is the

Eulerianization of Pρ0 and it is not difficult to write (199) in terms of this quantity.

38

Page 39: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Upon adopting this usual vector notation, the bracket (192) can also be written as

{F,G}∗ = −

∫d3x

[∇δGδρ· Pρ

δFδv− ∇

δFδρ· Pρ

δGδv

+∇sρ·

(δFδsPρδGδv−δGδsPρδFδv

)+∇ × vρ·

(PρδGδv× Pρ

δFδv

)+∇ · vρ

(δFδv· Pρ

δGδv−δGδv· Pρ

δFδv

)]. (199)

For MHD there is a magnetic field contribution to (191) and following the steps that lead to (199)

we obtain

{F,G}∗B = −

∫d3x

[B ·

(1ρPρδFδv· ∇

δGδB−

1ρPρδGδv· ∇

δFδB

)+ B ·

(∇

(1ρPρδFδv

)·δGδB− ∇

(1ρPρδGδv

)·δFδB

) ]. (200)

With the exception of the last term of (199) proportional to ∇ · v and the presence of the Eulerian

projection operator Pρ, (199) added to (200) is identical to the noncanonical Poisson bracket for

the ideal fluid and MHD as given in Morrison & Greene [74]. By construction, we know that (199)

satisfies the Jacobi identity – this follows because it was obtained by Eulerianizing the canonical

Dirac bracket in terms of Lagrangian variables. Guessing the bracket and proving Jacobi for (199)

directly would be a difficult chore, giving credence to the path we have followed in obtaining it.

To summarize, the bracket of (199) together with the Hamiltonian

H =12

∫d3x ρ |v|2 , (201)

the Eulerian counterpart of (187), generates dynamics that can preserve the constraint ∇ · v = 0. If

we add HB =∫

d3x |B|2/2 to (201) and add (200) to (199), then we obtain incompressible MHD.

The fluid case is the Eulerian counterpart of the volume preserving geodesic flow, described orig-

inally by Lagrange in Lagrange variables. Upon performing a series of straightforward manipula-

39

Page 40: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

tions, we obtain the following equations of motion for the flow:

∂ρ

∂t= {ρ,H}∗ = −∇ · Pρ

δHδv

= −∇ρ · Pρv , (202)

∂s∂t

= {s,H}∗ = −∇sρ· Pρ

δHδv

= −∇s · Pρv , (203)

∂v∂t

= {v,H}∗ = −1ρPρ

(ρ∇

δHδρ

)+

1ρPρ

(∇sδHδs

)−

1ρPρ

((∇ × v) × Pρ

δHδv

)−∇ · vρPρδHδv

+1ρPρ

(∇ · v

δHδv

)= Pρ∇

|v|2

2− Pρ

((∇ × v) × Pρv

)− (∇ · v)Pρv + Pρ (v∇ · v) . (204)

If we include HB we obtain additional terms to (204) generated by (200) for the projected J × B

force. Observe, equation (204) is not yet evaluated on the constraint D2 = 0, which in Eulerian

variables is ∇ · v = 0. As noted at the end of Section IV B, we turn to this task in Section IV D.

2. Eulerian-Dirac constraint theory direct with uniform density

For completness we recall the simpler case where the Eulerian density ρ is uniformly constant,

which without loss of generality can be scaled to unity. This case was considered in Chandre et al.

[55, 56], Nguyen & Turski [77, 78] (although a trick of using entropy as density was employed in

Chandre et al. [55] to treat density advection). In these works the Dirac constraints were chosen

to be the pointwise Eulerian quantities

D1 = ρ and D2 = ∇ · v , (205)

and the Dirac procedure was effected on the purely Eulerian level. This led to the projector

P := Pρ=1 = 1 − ∇∆−1∇ · , (206)

where ∆ = ∆ρ=1, and the following Dirac bracket:

{F,G}∗ = −

∫d3x

[∇sρ·

(δFδsPδGδv−δGδsPδFδv

)−∇ × vρ·

(PδFδv× P

δGδv

) ]. (207)

Incompressible MHD with constant density is generated by adding the following to (207)

{F,G}∗B = −

∫d3x

[Bρ·

(PδFδv· ∇

δGδB− P

δGδv· ∇

δFδB

)+B ·

(∇

(1ρPδFδv

)·δGδB− ∇

(1ρPδGδv

)·δFδB

) ], (208)

40

Page 41: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

and adding |B|2/2 to the integrand of (201).

The bracket of (207) differs from that of (199) in two ways: the projector Pρ is replaced by the

simpler projector P and it is missing the term proportional to ∇ ·v. Given that ∇ ·v cannot be set to

zero until after the equations of motion are obtained, this term gives rise to a significant differences

between the constant and nonconstant density Poisson brackets and incompressible dynamics.

D. Comparison of the Eulerian-Dirac and Lagrangian-Dirac constrained theories

Let us now discuss equations (202), (203) and (204). Given that ∇·Pρv = 0 (cf. (198)) it is clear

that the density and entropy are advected by the incompressible velocity field Pρv, as expected.

However, the meaning of (204) remains to be clarified. To this end we take the divergence of (204)

and again use (198) to obtain

∂(∇ · v)∂t

= −∇ ·(∇ · vPρv

)= −

(Pρv

)·∇ (∇ · v) . (209)

Thus ∇ · v itself is advected by an incompressible velocity field. As with any advection equation,

if initially ∇ · v = 0 , it will remain uniformly zero. After setting ∇ · v = 0 in equation (204) it

collapses down to∂v∂t

= −Pρ (v · ∇v) ; (210)

this is the anticipated equation of motion, the momentum equation of (1) with the insertion of the

pressure given by (5).

Given the discussion of Lagrangian vs. Eulerian constants of motion of Section (III D), that ∇·v

is advected rather than pointwise conserved is to be expected. Our development began with the

constraints D1,2 of (141) and (142) both of which are pointwise conserved by the Dirac procedure,

i.e. DL ≡ 0. This means their corresponding fluxes are identically zero, i.e., in (123) we have

ΓDL ≡ 0 for each. Thus the flux component ΓDE of (127) vanishes and the Eulerian flux for both

D1 and D2 have the form vDE. Because D1 and D2 Eulerianize to − ln(ρ) and ∇ · v, respectively,

we expected equations of the from of (209) for both. We will see in Section IV E that the equation

for D1 in fact follows also because the constraints are Casimir invariants.

Let us return to (190) and compare with the results of Section II C 1. Because the incompress-

ibility condition is an holonomic constraint and Section II C 1 concerns holonomic constraints for

the uncoupled N-body problem, both results are geodesic flows. In fact, one can think of the fluid

case as a continuum version of that of Section II C 1 with an infinity of holonomic constraints– thus

41

Page 42: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

we expect similarities between these results. However, because the incompressibility constraints

are pointwise constraints, the comparison is not as straightforward as it would be for global con-

straints of the fluid.

To make the comparison we first observe that the termAmn of (184) must correspond to the term↔

Ai j of (54), since their origin follows an analogous path in the derivation, both are antisymmetric,

and both project from both the left and the right. The analog of (56) according to (171) is

(Pρ0⊥

)i

j

π j

ρ0=ηi`

ρ0Au`

∂au

∆−1ρ0

J

Ahj

J

∂ah

π j

ρ0

=ηi`

ρ0Au`

∂au

∆−1ρ0

J

(D2

) ≡ 0 , (211)

when evaluated on D2 = 0. Unlike (56) a sum, which would here be an integral over d3a, does

not occur because the constraint D2 is a pointwise constraint as opposed to a global constraint.

Also, because the constraints are pointwise, the↔

Ti j is analogous to the terms with Tmn that also

have a factor of the projector Pρ0⊥, giving the results analogous to (57). Just as in Section II C 1,

we obtain πi = ρ0qi from (189) when evaluated on the constraint D2 = 0 and only a single term

involving the Tmn contributes to the momentum equation of motion (190) when evaluated on the

constraint D2 = 0, i.e.

πi = ρ0ηin

(Pρ0⊥

)n

r

(qm ηrsTms

)= Au

i∂

∂au

∆−1ρ0

J

(Ah`

J

∂ah

(qm Ak

m

J

∂q`

∂ak

))= Au

i∂

∂au

∆−1ρ0

J

(Ah`

J

∂ah

∂ak

(qm Ak

m

Jq`

))= Au

i∂

∂au

∆−1ρ0

J

(1J

∂ah Ah`A

km∂

∂ak

(qm q`

J

)) . (212)

Thus

qi = ηi` Au`

ρ0

∂au

∆−1ρ0

J

(1J

∂ah Ahj A

fk

∂a f

(q j qk

J

)) = − Γ ijk(q

j, qk) , (213)

where we have swapped indices and defined Γ ijk(q

j, qk), the normal force operator for geodesic

flow, analogous to that of (62). Like Γi, jk of (63), Γ ijk possesses symmetry: given arbitrary vector

fields V and W

Γ ijk(V

j,Wk) := −ηi` Au`

ρ0

∂au

∆−1ρ0

J

(1J

∂ah Ahj A

fk

∂a f

(V j Wk

J

)) = Γ ijk(V

k,W j) . (214)

Equation (213) defines geodesic flow on the group of volume preserving diffeomorphisms,

as was the case in Section II C 1, it does so in terms of the original coordinates, i.e., without

42

Page 43: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

specifically transforming to normal coordinates on the constraint surfaces which here are infintie-

dimensional.

Now we are in position to close the circle by writing (213) in Eulerian form. We will do this for

the ideal fluid, but MHD follows similarly. The term qi becomes as usual the advective derivative

∂v/∂t + v · ∇v and the Γijk term gives Pρ⊥ (v · ∇v). Thus (213) is precisely the Lagrangian form of

(210), written as follows:

∂v∂t

= −Pρ (v · ∇v) = −Pρ(∇ · (v ⊗ v)

). (215)

Similarly, the Lagrangian version of (209) follows easily from (213). To see this we operate with

the counterpart of taking the Eulerian divergence on the first line of (212) and make use of (182),

Ahn

J

∂ah

πn

ρ0=

Ahn

J

∂ah

(Pρ0⊥

)n

r

(qm ηrsTms

)=

Ahn

J

∂ah

(qm ηnsTms

)= δn

`

Ahn

J

∂ah

(qm Ak

m

J

∂q`

∂ak

), (216)

which in Eulerian variables becomes

∇ ·

(∂v∂t

+ v · ∇v)

= ∇ · (v · ∇v) or ∇ ·∂v∂t

=∂

∂t∇ · v = 0 . (217)

In Lagrangian variables we have the trivial conservation laws

ρ0 = 0 and s0 = 0 , (218)

where the corresponding fluxes are identically zero. However, as is evident from (202) and (203)

we obtain nontrivial conservation laws for ρ and s with nonzero fluxes. Thus we see again, con-

sistent with Section III D, how Lagrangian and Eulerian conservation laws are not equivalent.

For the special case where ρ0 = J = 1 one could proceed directly from (7), write it in terms of

the Lagrangian variables, and obtain (213). However, without the constraint theory, one would not

immediately see it is Hamiltonian and in fact geodesic flow on an infinite-dimensional manifold.

E. Incompressible algebra of invariants

In closing this section, we examine the constants of motion for the constrained system. The

Poisson bracket together with the set of functionals that commute with the Hamiltonian, i.e., that

satisfy {H, Ia} = 0 for a = 1, 2, . . . , d, constitute the d-dimensional algebra of invariants, a subal-

gebra of the infinite-dimensional Poisson bracket realization on all functionals. This subalgebra

43

Page 44: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

is a Lie algebra realization associated with a symmetry group of the dynamical system, and the

Poisson bracket with {Ia, · } yields the infinitesimal generators of the symmetries, i.e., the differ-

ential operator realization of the algebra. This was shown for compressible MHD in [71], where

the associated Lie algebra realization of the 10 parameter Galilean group on functionals was de-

scribed. This algebra is homomorphic to usual representations of the Galilean group, with the

Casimir invariants being in the center of the algebra composed of elements that have vanishing

Poisson bracket with all other elements.

A natural question to ask is what happens to this algebra when incompressibility is enforced by

our Dirac constraint procedure. Obviously the Hamiltonian is in the subalgebra and {H, · }∗ clearly

generates time translation, and this will be true for any Hamiltonian, but here we use Hamiltonian

of (201).

Inserting the momentum

P =

∫d3x ρv (219)

into (199) with the Hamiltonian (201) gives

{P,H}∗ = 0 (220)

without assuming ∇ · v = 0. To see this, we use (196) to obtain

PρδHδv

= ρv − ∇∆−1ρ ∇ · v and Pρ

δPi

δ3 j= ρ δi j , (221)

which when inserted into (199) gives

{Pi,H}∗ = −

∫d3x

2∂|v|2

∂xi+ 3i∇ ·

(ρv − ∇∆−1

ρ ∇ · v)

+[(∇ × v) ×

(ρv − ∇∆−1

ρ ∇ · v)]

i

+(∇ · v)[(ρv − ∇∆−1

ρ ∇ · v)

i− ρ3i

] ]= 0 , (222)

as expected. The result of (222) follows upon using standard vector identities, integration by parts,

and the self-adjointness of ∆−1ρ .

The associated generator of space translations that satisfies the constraints is given by the op-

erator {P, · }∗, which can be shown directly. And, it follows that

{Pi, P j}∗ = 0 , ∀ i, j = 1, 2, 3 . (223)

Because the momentum contains no s dependence the the second line of (199) vanishes and using

PρδPi/δ3 j = ρ δi j of (221) it is clear the last line involving ∇ · v of (199) also vanishes. The result

of (223) is obtained because the first and third lines cancel.

44

Page 45: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Next, consider the angular momentum

L =

∫d3x ρ x × v . (224)

We will show

{Li,H}∗ = 0 . (225)

Using PρδLi/δv = δLi/δv, the fact that {Li,H} = 0 for the compressible fluid, and Pρ = I − Pρ⊥,

we obtain

{Li,H}∗ =

∫d3x

[− ∇

δLi

δρ· Pρ⊥(ρv)

+

(δLi

δv×∇ × vρ

+∇ · vρ

δLi

δv

)· Pρ⊥(ρv)

]. (226)

Next, recognizing that Pρ⊥(ρv) = ∇∆−1ρ ∇ · v and integrating by parts, we obtain

{Li,H}∗ =

∫d3x ∆−1

ρ (∇ · v)[∇2 δLi

δρ− ∇ ·

(δLi

δv×∇ × vρ

+δLi

δv∇ · vρ

) ]. (227)

Then upon insertingδLi

δρ= εi jkx j3k and

δLi

δ3 j= ρ xkεik j , (228)

and using standard vector analysis we obtain (225).

Because PρδLi/δv = δLi/δv, the first and third lines of (199) produce

{Li, L j}∗ = εi jkLk , (229)

just as they do for the compressible fluid (and MHD), while the fourth line manifestly vanishes.

Similarly, it follows that that {L, · }∗ is the generator for rotations.

To obtain the full algebra of invariants we need {Li, P j}∗. However because PρδPi/δv = δPi/δv

and PρδLi/δv = δLi/δv, it follows as for the compressible fluid that {Li, P j}∗ = εi jkPk.

Finally, consider the following measure of the position of the center of mass, the generator of

Galilean boosts,

G =

∫d3x ρ (x − vt) . (230)

Calculations akin to those above reveal

{Gi,G j}∗ = 0 , {Gi, P j}∗ = 0 , {Gi,H}∗ = Pi , {Li,G j}∗ = εi jkGk . (231)

Thus the bracket (207) with the set of ten invariants {H,P,L,G} is at once a closed subalge-

bra of Poisson bracket realization on all functionals and produces an operator realization of the

45

Page 46: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Galilean group [see e.g. 82] that is homomorphic to the operator algebra of {Li, · }∗, {Pi, · }∗, etc.

with operator commutation relations. This remains true for MHD with the only change being the

addition of HB to the Hamiltonian.

Thus, the Galilean symmetry properties of the ideal fluid and MHD are not affected by the

compressibility constraint. However, based on past experience with advected quantities, we do

expect a new Casimir invariant of the form

C[ρ, s] =

∫d3x C(ρ, s) . (232)

To see that {C, F}∗ = 0 for any functional F, where C(ρ, s) is an arbitrary function of its arguments,

we calculate

{F, C}∗ = −

∫d3x

[ρ∇

∂C

∂ρ−∂C

∂s∇s

]· Pρ

δFδv

, (233)

and since ∇ × (ρ∇∂C/∂ρ − ∂C/∂s∇s) = 0 we write it as ∇p, giving for (233)

{F, C}∗ = −

∫d3x

1ρ∇p · Pρ

δFδv

. (234)

Thus, integration by parts and use of (198) imply {F, C}∗ = 0 for all functionals F. Note, without

loss of generality we can write C(ρ, s) = ρU(ρ, s), in which case p = ρ2∂U/∂ρ. Thus, it is

immaterial whether or not one retains the internal energy term∫

d3x ρU(ρ, s) in the Hamiltonian.

Now, (232) is not the most general Casimir. Because both ρ and ∇ · v are Lagrangian pointwise

Dirac constraints, we expect the following to be an Eulerian Casimir

C[ρ, s,∇ · v] =

∫d3xC(ρ, s,∇ · v) , (235)

where C is an arbitrary function of its arguments. To see that {C, F}∗ = 0 for any functional F, we

first observe thatδCδv

= −∇∂C

∂∇ · v(236)

and, as is evident from (196), that ∇ · (Pρ∇Φ) = 0 for all Φ; hence, all the δC/δv terms vanish

except the first term of the last line of (199). This term combines with the others to cancel, just as

for the calculation of C.

For constant density, entropy, and magnetic field, the bracket of (207) reduces to

{F,G}∗ = −

∫d3x∇ × vρ·

(PδFδv× P

δGδv

), (237)

whence it is easily seen that the helicity

C3·∇×3 =

∫d3x v · ∇ × v (238)

46

Page 47: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

is a Casimir invariant because P (∇ × v) = ∇ × v. This Casimir is lost when entropy and density

are allowed to be advected, for it is no longer a Casimir invariant of (199).

Now, let us consider invariants in the Lagrangian description. Without the incompressibility

constraints, the Hamiltonian has a standard kinetic energy term and the internal energy depends

on ∂q/∂a, an infinitesimal version of the two-body interaction, if follows that just like the N-body

problem the system has Galilean symmetry, and because the Poisson bracket in the Lagrangian de-

scription (100) is canonical there are no Casimir invariants. With the incompressibility constraint,

the generators of the algebra now respect the constraints, with Dirac constraints being Casimirs

and the algebra of constraints now having a nontrivial center. Because the Casimirs are pointwise

invariants, we expect the situation to be like that for the Maxwell Vlasov equation [71], where the

following is a Casimir

C∇·B[B] =

∫d3xC(∇ · B, x) , (239)

where here C is an arbitrary function of its arguments. This suggests the following:

C[J] =

∫d3a C(J, a) . (240)

Indeed, only the first term of (184) contributes when we calculate {C,G}∗ and this term vanishes

by (181) becauseδCδqi = −

∂a`

(A`

i∂C

∂J

). (241)

Similarly, it can be shown that the full Casimir is

C[D1,D2] =

∫d3a C(D1,D2, a) , (242)

a Lagrangian Casimir consistent with (235).

For MHD, the magnetic helicity,

CA·B =

∫d3x A · B , (243)

where B = ∇ × A is easily seen to be preserved and a Casimir up to the usual issues regarding

gauge conditions and boundary terms [see 64]. We know that the cross helicity

C3·B =

∫d3x v · B , (244)

is a Casimir of the compressible barotropic MHD equations, and it is easy to verify that it is also

a Casimir of (207) added to (208), that is for uniform density. However, it is not a Casimir for the

case with advected density, i.e., for the bracket of (199) added to (200).

47

Page 48: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

V. CONCLUSIONS

In this paper we have substantially investigated constraints, particularly incompressibility for

the ideal fluid and MHD, for the three dichotomies described in Section I A: the Lagrangian vs.

Eulerian fluid descriptions, Lagrange multiplier vs. Dirac constraint methods, and Lagrangian vs.

Hamiltonian formalisms. An in depth description of the interplay between the various fluid and

MHD descriptions was given, with an emphasis on Dirac’s constraint method. Although we mainly

considered geodesic flow for simplicity, the Dirac’s Poisson bracket method can be used to find

other forces of constraint in a variety of fluid and plasma contexts.

Based on our results, many avenues for future research are presented. We mention a few. Since

the Hamiltonian structure of extended and relativistic MHD are now at hand [44, 58, 60, 61, 67, 70]

calculations analogous to those presented here can be done for a variety of magnetofluid models.

Another avenue for future research would be to address stability with constraints. In a previous

series of papers [45, 46, 47, 48, 49] we have investigated Hamiltonian based stability, generaliza-

tions of the MHD energy principle or the ideal fluid Rayleigh criterion, within the Lagrangian,

energy-Casimir, and dynamically accessible frameworks. Because Dirac’s method adds Casimirs,

the Dirac constraints, one gets a richer set of equilibria from the energy-Casimir variational princi-

ple and these can be tested for Lyapunov stability. Similarly, the method of dynamical accessibility

[see 72] based on constrained variations induced by the Poisson operator will enlarge the set of

stable equilibria.

Lastly, we mention that there is considerable geometric structure behind our calculations. These

can be restated in geometric/Lie group language [see e.g. 54]. Also, Arnold’s program for obtain-

ing the Riemann curvature for geodesic flow on the group of volume preserving diffeomorphisms

can be explored beginning from our results of Section IV. We did not feel this special issue would

be the appropriate place to explore these ideas.

ACKNOWLEDGMENT

PJM was supported by U.S. Dept. of Energy under contract #DE-FG02-04ER-54742. He would

also like to acknowledge support from the Humboldt Foundation and the hospitality of the Numer-

ical Plasma Physics Division of the IPP, Max Planck, Garching. FP would like to acknowledge the

48

Page 49: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

hospitality of the Institute for Fusion Studies of the University of Texas at Austin.

[44] Abdelhamid, H. M., Kawazura, Y. & Yoshida, Z. 2015 Hamiltonian formalism of extended magneto-

hydrodynamics. J. Phys. A 48 (23), 235502.

[45] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2010 MHD equilibrium variational principles with

symmetry. Plasma Phys. and Controll. Fusion P52, 055001.

[46] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2012 Hamiltonian magnetohydrodynamics: Symmetric

formulation, Casimir invariants, and equilibrium variational principles. Phys. Plasmas 19, 052102.

[47] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2013 Hamiltonian magnetohydrodynamics: La-

grangian, Eulerian, and dynamically accessible stability - Theory. Phys. Plasmas 20, 092104.

[48] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2015 Erratum: Hamiltonian magnetohydrodynamics:

Lagrangian, Eulerian, and dynamically accessible stability - Theory. Phys. Plasmas 22, 039903.

[49] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2016 Hamiltonian magnetohydrodynamics: La-

grangian, Eulerian, and dynamically accessible stability - examples with translation symmetry. Phys.

Plasmas 23, 102112.

[50] Arnold, V. I. 1966 Sur la geometrie differentielle des groupes de Lie de dimension infinie et ses

applications a l’hydrodynamique des fluides parfaits. Ann. l’Inst. Fourier XVI, 319–361.

[51] Arnold, V. I. 1978 Mathematical Methods of Classical Mechanics. Springer-Verlag.

[52] Arnold, V. I. & Khesin, B. A. 1998 Topological Methods in Hydrodynamics. Springer-Verlag.

[53] Arnold, V. I., Kozlov, V. V. & Neishtadt, A. I. 1980 Dynamical Systems III. Springer Verlag, New

York.

[54] Bloch, A. M. 2002 Nonholonomic Mechanics and Control. Springer-Verlag.

[55] Chandre, C., de Guillebon, L., Back, A., Tassi, E. & Morrison, P. J. 2013 On the use of projectors

for Hamiltonian systems and their relationship with Dirac brackets. J. Phys. A 46, 125203.

[56] Chandre, C., Morrison, P. J. & Tassi, E. 2012 On the Hamiltonian formulation of incompressible

ideal fluids and magnetohydrodynamics via Dirac’s theory of constraints. Phys. Lett. A 376, 737–743.

[57] Chandre, C., Morrison, P. J. & Tassi, E. 2014 Hamiltonian formulation of the modified Hasegawa-

Mima equation. Phys. Lett A 378, 956–959.

[58] Charidakos, I. Keramidas, Lingam, M., Morrison, P. J., White, R. L. & Wurm, A. 2014 Action

principles for extended MHD models. Phys. Plasmas 21, 092118.

49

Page 50: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

[59] Corben, H. C. & Stehle, P. 1960 Classical Mechanics, 2nd edn. Wiley.

[60] D’Avignon, E., Morrison, P. J. & Pegoraro, F. 2015 Action principle for relativistic magnetohydro-

dynamics. Phys. Rev. D 91, 084050.

[61] D’Avignon, E. C., Morrison, P. J. & Lingam, M. 2016 Derivation of the Hall and extended magneto-

hydrodynamics brackets. Phys. Plasmas 23, 062101.

[62] Dirac, P. A. M. 1950 Generalized Hamiltonian dynamics. Canadian J. Math. 2, 129–148.

[63] Evans, L. C. 2010 Partial Differential Equations. American Mathematical Society, Providence, RI.

[64] Finn, J. M. & Antonsen, T. M. 1985 Magnetic helicity: What is it and what is it good for? Comments

Plasma Phys. Control. Fusion 9, 111–126.

[65] Flierl, G. R. & Morrison, P. J. 2011 Hamiltonian-Dirac simulated annealing: Application to the

calculation of vortex states. Physica D 240, 212–232.

[66] Hanson, A., Regge, T. & Teitleboim, C. 1976 Constrained Hamiltonian Systems. Accademia

Nazionale dei Lincei, Rome.

[67] Kaltsas, D. A., Throumoulopoulos, G. N. & Morrison, P. J. 2020 Energy-Casimir, dynamically

accessible, and Lagrangian stability of extended magnetohydrodynamic equilibria. Phys. Plasmas 27,

012104.

[68] Kampen, N. G. Van& Felderhoff, B. U. 1967 Theoretical Methods in Plasma Physics. North-Holland,

Amsterdam.

[69] Lagrange, J. L. 1788 Mecanique Analytique. Ve Courcier, Paris.

[70] Lingam, M., Miloshevich, G. & Morrison, P. J. 2016 Concomitant Hamiltonian and topological struc-

tures of extended magnetohydrodynamics. Phys. Lett. A 380, 2400–2406.

[71] Morrison, P. J. 1982 Poisson brackets for fluids and plasmas. AIP Conf. Proc. 88, 13–46.

[72] Morrison, P. J. 1998 Hamiltonian description of the ideal fluid. Rev. Mod. Phys. 70, 467–521.

[73] Morrison, P. J. 2009 On Hamiltonian and action principle formulations of plasma dynamics. AIP

Conf. Proc. 1188, 329–344.

[74] Morrison, P. J. & Greene, J. M. 1980 Noncanonical Hamiltonian density formulation of hydrody-

namics and ideal magnetohydrodynamics. Phys. Rev. Lett. 45, 790–793.

[75] Morrison, P. J., Lebovitz, N. R. & Biello, J. 2009 The Hamiltonian description of incompressible

fluid ellipsoids. Ann. Phys. 324, 1747–1762.

[76] Newcomb, WA 1962 Lagrangian and Hamiltonian methods in magnetohydrodynamics. Nuclear Fu-

sion Supp. 2, 451–463.

50

Page 51: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

[77] Nguyen, S. & Turski, L. A. 1999 Canonical description of incompressible fluid: Dirac brackets ap-

proach. Physica A 272, 48–55.

[78] Nguyen, S. & Turski, L. A. 2001 Examples of the Dirac approach to dynamics of systems with

constraints. Physica A 290, 431–44.

[79] Orszag, S. A., Israeli, M. & Deville, M. O. 1986 Boundary conditions for incompressible flows. J.

Sci. Comp. 1, 75–111.

[80] Serrin, J. 1959 Mathematical principles of classical fluid mechanics. In Handbuch der Physik VIII

pt. 1 (ed. S. Flugge), pp. 125–262. Springer Verlag.

[81] Sommerfeld, A. 1964 Mechanics of Deformable Bodies. Academic Press.

[82] Sudarshan, E. C. G. & Makunda, N. 1974 Classical Dynamics : A Modern Perspective. John Wiley

& Sons, New York.

[83] Sundermeyer, K. 1982 Constrained Dynamics, Lecture Notes in Physics vol. 169. Springer Verlag,

New York.

[84] Tassi, E., Chandre, D. & Morrison, P. J. 2009 Hamiltonian derivation of the Charney-Hasegawa-

Mima equation. Phys. Plasmas 16, 082301.

[85] Van Kampen, N. G. & Felderhof, B. U. 1967 Theoretical Methods in Plasma Physics. North Holland,

Amsterdam.

[86] Whittaker, E. T. 1917 A Treatise on the Analytical Dynamics of Particles and Rigid Bodies, 2nd edn.

Cambridge University Press.

[44] Abdelhamid, H. M., Kawazura, Y. & Yoshida, Z. 2015 Hamiltonian formalism of extended magneto-

hydrodynamics. J. Phys. A 48 (23), 235502.

[45] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2010 MHD equilibrium variational principles with

symmetry. Plasma Phys. and Controll. Fusion P52, 055001.

[46] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2012 Hamiltonian magnetohydrodynamics: Symmetric

formulation, Casimir invariants, and equilibrium variational principles. Phys. Plasmas 19, 052102.

[47] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2013 Hamiltonian magnetohydrodynamics: La-

grangian, Eulerian, and dynamically accessible stability - Theory. Phys. Plasmas 20, 092104.

[48] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2015 Erratum: Hamiltonian magnetohydrodynamics:

Lagrangian, Eulerian, and dynamically accessible stability - Theory. Phys. Plasmas 22, 039903.

[49] Andreussi, T., Morrison, P. J. & Pegoraro, F. 2016 Hamiltonian magnetohydrodynamics: La-

grangian, Eulerian, and dynamically accessible stability - examples with translation symmetry. Phys.

51

Page 52: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

Plasmas 23, 102112.

[50] Arnold, V. I. 1966 Sur la geometrie differentielle des groupes de Lie de dimension infinie et ses

applications a l’hydrodynamique des fluides parfaits. Ann. l’Inst. Fourier XVI, 319–361.

[51] Arnold, V. I. 1978 Mathematical Methods of Classical Mechanics. Springer-Verlag.

[52] Arnold, V. I. & Khesin, B. A. 1998 Topological Methods in Hydrodynamics. Springer-Verlag.

[53] Arnold, V. I., Kozlov, V. V. & Neishtadt, A. I. 1980 Dynamical Systems III. Springer Verlag, New

York.

[54] Bloch, A. M. 2002 Nonholonomic Mechanics and Control. Springer-Verlag.

[55] Chandre, C., de Guillebon, L., Back, A., Tassi, E. & Morrison, P. J. 2013 On the use of projectors

for Hamiltonian systems and their relationship with Dirac brackets. J. Phys. A 46, 125203.

[56] Chandre, C., Morrison, P. J. & Tassi, E. 2012 On the Hamiltonian formulation of incompressible

ideal fluids and magnetohydrodynamics via Dirac’s theory of constraints. Phys. Lett. A 376, 737–743.

[57] Chandre, C., Morrison, P. J. & Tassi, E. 2014 Hamiltonian formulation of the modified Hasegawa-

Mima equation. Phys. Lett A 378, 956–959.

[58] Charidakos, I. Keramidas, Lingam, M., Morrison, P. J., White, R. L. & Wurm, A. 2014 Action

principles for extended MHD models. Phys. Plasmas 21, 092118.

[59] Corben, H. C. & Stehle, P. 1960 Classical Mechanics, 2nd edn. Wiley.

[60] D’Avignon, E., Morrison, P. J. & Pegoraro, F. 2015 Action principle for relativistic magnetohydro-

dynamics. Phys. Rev. D 91, 084050.

[61] D’Avignon, E. C., Morrison, P. J. & Lingam, M. 2016 Derivation of the Hall and extended magneto-

hydrodynamics brackets. Phys. Plasmas 23, 062101.

[62] Dirac, P. A. M. 1950 Generalized Hamiltonian dynamics. Canadian J. Math. 2, 129–148.

[63] Evans, L. C. 2010 Partial Differential Equations. American Mathematical Society, Providence, RI.

[64] Finn, J. M. & Antonsen, T. M. 1985 Magnetic helicity: What is it and what is it good for? Comments

Plasma Phys. Control. Fusion 9, 111–126.

[65] Flierl, G. R. & Morrison, P. J. 2011 Hamiltonian-Dirac simulated annealing: Application to the

calculation of vortex states. Physica D 240, 212–232.

[66] Hanson, A., Regge, T. & Teitleboim, C. 1976 Constrained Hamiltonian Systems. Accademia

Nazionale dei Lincei, Rome.

[67] Kaltsas, D. A., Throumoulopoulos, G. N. & Morrison, P. J. 2020 Energy-Casimir, dynamically

accessible, and Lagrangian stability of extended magnetohydrodynamic equilibria. Phys. Plasmas 27,

52

Page 53: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

012104.

[68] Kampen, N. G. Van& Felderhoff, B. U. 1967 Theoretical Methods in Plasma Physics. North-Holland,

Amsterdam.

[69] Lagrange, J. L. 1788 Mecanique Analytique. Ve Courcier, Paris.

[70] Lingam, M., Miloshevich, G. & Morrison, P. J. 2016 Concomitant Hamiltonian and topological struc-

tures of extended magnetohydrodynamics. Phys. Lett. A 380, 2400–2406.

[71] Morrison, P. J. 1982 Poisson brackets for fluids and plasmas. AIP Conf. Proc. 88, 13–46.

[72] Morrison, P. J. 1998 Hamiltonian description of the ideal fluid. Rev. Mod. Phys. 70, 467–521.

[73] Morrison, P. J. 2009 On Hamiltonian and action principle formulations of plasma dynamics. AIP

Conf. Proc. 1188, 329–344.

[74] Morrison, P. J. & Greene, J. M. 1980 Noncanonical Hamiltonian density formulation of hydrody-

namics and ideal magnetohydrodynamics. Phys. Rev. Lett. 45, 790–793.

[75] Morrison, P. J., Lebovitz, N. R. & Biello, J. 2009 The Hamiltonian description of incompressible

fluid ellipsoids. Ann. Phys. 324, 1747–1762.

[76] Newcomb, WA 1962 Lagrangian and Hamiltonian methods in magnetohydrodynamics. Nuclear Fu-

sion Supp. 2, 451–463.

[77] Nguyen, S. & Turski, L. A. 1999 Canonical description of incompressible fluid: Dirac brackets ap-

proach. Physica A 272, 48–55.

[78] Nguyen, S. & Turski, L. A. 2001 Examples of the Dirac approach to dynamics of systems with

constraints. Physica A 290, 431–44.

[79] Orszag, S. A., Israeli, M. & Deville, M. O. 1986 Boundary conditions for incompressible flows. J.

Sci. Comp. 1, 75–111.

[80] Serrin, J. 1959 Mathematical principles of classical fluid mechanics. In Handbuch der Physik VIII

pt. 1 (ed. S. Flugge), pp. 125–262. Springer Verlag.

[81] Sommerfeld, A. 1964 Mechanics of Deformable Bodies. Academic Press.

[82] Sudarshan, E. C. G. & Makunda, N. 1974 Classical Dynamics : A Modern Perspective. John Wiley

& Sons, New York.

[83] Sundermeyer, K. 1982 Constrained Dynamics, Lecture Notes in Physics vol. 169. Springer Verlag,

New York.

[84] Tassi, E., Chandre, D. & Morrison, P. J. 2009 Hamiltonian derivation of the Charney-Hasegawa-

Mima equation. Phys. Plasmas 16, 082301.

53

Page 54: Lagrangian and Dirac constraints for the ideal ...morrison/20_MAPIIc_ArX.pdfSITAEL S.p.A., Pisa, 56121, Italy F. Pegoraro Dipartimento di Fisica E. Fermi, Pisa, 56127, Italy Abstract

[85] Van Kampen, N. G. & Felderhof, B. U. 1967 Theoretical Methods in Plasma Physics. North Holland,

Amsterdam.

[86] Whittaker, E. T. 1917 A Treatise on the Analytical Dynamics of Particles and Rigid Bodies, 2nd edn.

Cambridge University Press.

54